243
Copyright by Arash Fathi 2015

Diss_AF

Embed Size (px)

Citation preview

Copyright

by

Arash Fathi

2015

The Dissertation Committee for Arash Fathicertifies that this is the approved version of the following dissertation:

Full-waveform inversion in three-dimensional

PML-truncated elastic media:

theory, computations, and field experiments

Committee:

Loukas F. Kallivokas, Supervisor

Clinton N. Dawson

Leszek F. Demkowicz

Omar Ghattas

Lance Manuel

Kenneth H. Stokoe II

Full-waveform inversion in three-dimensional

PML-truncated elastic media:

theory, computations, and field experiments

by

Arash Fathi, B.Sc., M.Sc.

DISSERTATION

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

THE UNIVERSITY OF TEXAS AT AUSTIN

May 2015

I don’t tell the murky world

to turn pure.

I purify myself

and check my reflection

in the water of the valley brook.

Ryokan

Acknowledgments

This dissertation would not have been possible without the help of many oth-

ers. I would like to express my sincere gratitude and appreciation to my doctoral

advisor, Professor Loukas F. Kallivokas, for his assistance, ingenuity, and supervi-

sion of this work. I am thankful to Professors Clint Dawson, Leszek Demkowicz,

Omar Ghattas, Lance Manuel, and Kenneth Stokoe, for serving on my dissertation

committee, and for their valuable comments. I also wish to thank Babak Poursartip,

Georg Stadler, Sezgin Kucukcoban, Pranav Karve, and Chanseok Jeong, for their

continuous support, and fruitful discussions, and Farn-Yuh Menq, Y.-C. Lin and

Changyoung Kim, for assisting with the field experiments, processing of the SASW

test results, and the CPT tests, respectively.

Partial support for my research has been provided by the National Science

Foundation under grant award CMMI-0619078 and through an Academic Alliance

Excellence grant between the King Abdullah University of Science and Technology

in Saudi Arabia (KAUST) and the University of Texas at Austin. This support is

gratefully acknowledged.

v

Full-waveform inversion in three-dimensional

PML-truncated elastic media:

theory, computations, and field experiments

Arash Fathi, Ph.D.

The University of Texas at Austin, 2015

Supervisor: Loukas F. Kallivokas

We are concerned with the high-fidelity subsurface imaging of the soil, which

commonly arises in geotechnical site characterization and geophysical explorations.

Specifically, we attempt to image the spatial distribution of the Lame parameters in

semi-infinite, three-dimensional, arbitrarily heterogeneous formations, using surficial

measurements of the soil’s response to probing elastic waves. We use the com-

plete waveforms of the medium’s response to drive the inverse problem. Specifically,

we use a partial-differential-equation (PDE)-constrained optimization approach, di-

rectly in the time-domain, to minimize the misfit between the observed response of

the medium at select measurement locations, and a computed response correspond-

ing to a trial distribution of the Lame parameters. We discuss strategies that lend

algorithmic robustness to the proposed inversion schemes. To limit the computa-

tional domain to the size of interest, we employ perfectly-matched-layers (PMLs).

vi

The PML is a buffer zone that surrounds the domain of interest, and enforces the

decay of outgoing waves.

In order to resolve the forward problem, we present a hybrid finite element ap-

proach, where a displacement-stress formulation for the PML is coupled to a standard

displacement-only formulation for the interior domain, thus leading to a computa-

tionally cost-efficient scheme. We discuss several time-integration schemes, including

an explicit Runge-Kutta scheme, which is well-suited for large-scale problems on par-

allel computers.

We report numerical results demonstrating stability and efficacy of the for-

ward wave solver, and also provide examples attesting to the successful reconstruc-

tion of the two Lame parameters for both smooth and sharp profiles, using synthetic

records. We also report the details of two field experiments, whose records we subse-

quently used to drive the developed inversion algorithms in order to characterize the

sites where the field experiments took place. We contrast the full-waveform-based

inverted site profile against a profile obtained using the Spectral-Analysis-of-Surface-

Waves (SASW) method, in an attempt to compare our methodology against a widely

used concurrent inversion approach. We also compare the inverted profiles, at select

locations, with the results of independently performed, invasive, Cone Penetrometer

Tests (CPTs).

Overall, whether exercised by synthetic or by physical data, the full-waveform

inversion method we discuss herein appears quite promising for the robust subsur-

face imaging of near-surface deposits in support of geotechnical site characterization

investigations.

vii

Table of Contents

Acknowledgments v

Abstract vi

List of Tables xii

List of Figures xiii

Chapter 1. Introduction 1

1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.1.1 The perfectly-matched-layer (PML) . . . . . . . . . . . . . . . 3

1.1.2 Full-waveform inversion (FWI) . . . . . . . . . . . . . . . . . . 7

1.2 Present approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.3 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1.4 Dissertation outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

Chapter 2. Simulation of wave motion in three-dimensional PML-truncated heterogeneous media 15

2.1 Complex-coordinate-stretching . . . . . . . . . . . . . . . . . . . . . . 16

2.1.1 Key idea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.1.2 Choice of stretching functions . . . . . . . . . . . . . . . . . . . 18

2.2 Three-dimensional unsplit-field PML . . . . . . . . . . . . . . . . . . 20

2.2.1 Frequency-domain equations . . . . . . . . . . . . . . . . . . . 21

2.2.2 Time-domain equations . . . . . . . . . . . . . . . . . . . . . . 24

2.3 Hybrid finite element implementation . . . . . . . . . . . . . . . . . . 26

2.3.1 Spatial discretization . . . . . . . . . . . . . . . . . . . . . . . . 26

2.3.2 Discretization in time . . . . . . . . . . . . . . . . . . . . . . . 33

2.4 Spectral elements and explicit time integration . . . . . . . . . . . . . 35

2.5 A symmetric formulation . . . . . . . . . . . . . . . . . . . . . . . . . 37

viii

2.6 Generalization for multi-axial perfectly-matched-layers . . . . . . . . . 41

2.7 Numerical Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . 44

2.7.1 Homogeneous media . . . . . . . . . . . . . . . . . . . . . . . . 47

2.7.2 Heterogeneous media . . . . . . . . . . . . . . . . . . . . . . . 56

2.7.3 Comparison of various formulations . . . . . . . . . . . . . . . 63

2.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

Chapter 3. The elastic inverse medium problem in three-dimensionalPML-truncated domains 67

3.1 The forward problem . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

3.2 The inverse medium problem . . . . . . . . . . . . . . . . . . . . . . . 70

3.2.1 Optimality system . . . . . . . . . . . . . . . . . . . . . . . . . 73

3.2.1.1 The state problem . . . . . . . . . . . . . . . . . . . . 74

3.2.1.2 The adjoint problem . . . . . . . . . . . . . . . . . . . 74

3.2.1.3 The control problems . . . . . . . . . . . . . . . . . . . 77

3.2.2 The inversion process . . . . . . . . . . . . . . . . . . . . . . . 79

3.2.3 Buttressing schemes . . . . . . . . . . . . . . . . . . . . . . . . 81

3.2.3.1 Regularization factor selection and continuation . . . . 82

3.2.3.2 Source-frequency continuation . . . . . . . . . . . . . . 83

3.2.3.3 Biased search direction for λ . . . . . . . . . . . . . . . 84

3.3 Numerical experiments . . . . . . . . . . . . . . . . . . . . . . . . . . 85

3.3.1 Numerical verification of the material gradients . . . . . . . . . 87

3.3.2 Smoothly varying heterogeneous medium . . . . . . . . . . . . 90

3.3.2.1 Single parameter inversion . . . . . . . . . . . . . . . . 91

3.3.2.2 Simultaneous inversion . . . . . . . . . . . . . . . . . . 95

3.3.3 Layered medium . . . . . . . . . . . . . . . . . . . . . . . . . . 98

3.3.4 Layered medium with inclusion . . . . . . . . . . . . . . . . . . 104

3.3.5 Layered medium with three inclusions . . . . . . . . . . . . . . 109

3.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

ix

Chapter 4. Site characterization using full-waveform inversion 122

4.1 The inverse medium problem in two space dimensions . . . . . . . . . 123

4.1.1 The forward problem . . . . . . . . . . . . . . . . . . . . . . . 124

4.1.2 The inverse problem . . . . . . . . . . . . . . . . . . . . . . . . 130

4.2 The field experiment - design considerations . . . . . . . . . . . . . . 135

4.2.1 Line load truncation and spacing requirements . . . . . . . . . 138

4.2.2 Verification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

4.2.3 Signal design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

4.2.4 Parametric studies . . . . . . . . . . . . . . . . . . . . . . . . . 148

4.2.5 The experiment layout . . . . . . . . . . . . . . . . . . . . . . . 149

4.3 Field experiment records and data processing . . . . . . . . . . . . . . 152

4.3.1 Signal processing . . . . . . . . . . . . . . . . . . . . . . . . . . 153

4.3.2 Data integration . . . . . . . . . . . . . . . . . . . . . . . . . . 155

4.4 Inversion results using field experiment data . . . . . . . . . . . . . . 159

4.4.1 Inversion process . . . . . . . . . . . . . . . . . . . . . . . . . . 159

4.4.2 Comparison with SASW . . . . . . . . . . . . . . . . . . . . . . 161

4.4.3 Comparison with cone penetration test (CPT) results . . . . . 164

4.5 Three-dimensional site characterization . . . . . . . . . . . . . . . . . 166

4.5.1 The experiment site and layout . . . . . . . . . . . . . . . . . . 168

4.5.2 Pre-processing the field data . . . . . . . . . . . . . . . . . . . 170

4.5.3 Full-waveform inversion using field data . . . . . . . . . . . . . 170

4.5.4 Profiling obtained from SASW . . . . . . . . . . . . . . . . . . 178

4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

Chapter 5. Conclusions 186

5.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

5.2 Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

Appendices 190

x

Appendix A. Submatrix definitions 191

A.1 Submatrices in equation (2.31) . . . . . . . . . . . . . . . . . . . . . . 191

A.2 Submatrices for the symmetric PML formulation . . . . . . . . . . . . 192

A.3 Submatrices for M-PML . . . . . . . . . . . . . . . . . . . . . . . . . 193

A.4 Discretization of the control problems . . . . . . . . . . . . . . . . . . 193

A.5 Submatrices in equation (4.3) . . . . . . . . . . . . . . . . . . . . . . 194

Appendix B. Time-integration schemes 196

B.1 Fourth-order Runge-Kutta method . . . . . . . . . . . . . . . . . . . . 196

B.2 Extended Newmark method . . . . . . . . . . . . . . . . . . . . . . . 197

B.3 The adjoint problem time-integration scheme . . . . . . . . . . . . . . 199

Appendix C. Gradient of a functional 201

Appendix D. On the third discrete optimality condition 202

Appendix E. On the singular convolution integral (4.21) 204

Appendix F. On the spatial integration of (4.25) 209

Bibliography 212

Vita 226

xi

List of Tables

1.1 PML developments in time-domain elastodynamics. . . . . . . . . . . 6

2.1 Legendre-Gauss-Lobatto quadrature rule. . . . . . . . . . . . . . . . . 36

2.2 Discretization details of the hybrid-PML and enlarged domain models. 48

2.3 Relative error at sampling points between hybrid-PML and enlargeddomain solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.4 Test cases for comparing various formulations and their correspondingrelative error. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

3.1 Characterization of Gaussian pulses. . . . . . . . . . . . . . . . . . . 86

3.2 Comparison of the directional derivatives. . . . . . . . . . . . . . . . 90

4.1 Material properties used in load verification examples. . . . . . . . . . 141

4.2 Chirp signals used in the field experiment. . . . . . . . . . . . . . . . 147

xii

List of Figures

1.1 Problem definition: (a) interrogation of a heterogeneous semi-infinitedomain by an active source; and (b) computational model truncatedfrom the semi-infinite medium via the introduction of PMLs. . . . . . 10

2.1 A PML truncation boundary in the direction of coordinate s. . . . . . 17

2.2 PML-truncated semi-infinite domain. . . . . . . . . . . . . . . . . . . 28

2.3 Partitioning of submatrices in (2.31b). . . . . . . . . . . . . . . . . . 32

2.4 Ricker pulse time history and its Fourier spectrum. . . . . . . . . . . 45

2.5 PML-truncated semi-infinite homogeneous media. . . . . . . . . . . . 48

2.6 Snapshots of total displacement taken at t = 0.111 s, 0.219 s (verticalexcitation). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

2.7 Snapshots of total displacement taken at t = 0.147 s, 0.219 s (hori-zontal excitation). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.8 Comparison of displacement time histories between the enlarged andPML-truncated domain solutions at the sp2, sp4, sp6, and sp8 sam-pling points (homogeneous case, vertical excitation). . . . . . . . . . . 51

2.9 Comparison of displacement time histories between the enlarged andPML-truncated domain solutions at the sp2 and sp8 sampling points(homogeneous case, horizontal excitation). . . . . . . . . . . . . . . . 52

2.10 Relative error time history e(x, t) at various sampling points (homo-geneous case, vertical excitation). . . . . . . . . . . . . . . . . . . . . 53

2.11 Relative error time history e(x, t) at various sampling points (homo-geneous case, horizontal excitation). . . . . . . . . . . . . . . . . . . . 54

2.12 Total decay of energy within the regular domain for various values ofβo (homogeneous case, vertical excitation). . . . . . . . . . . . . . . . 55

2.13 Total decay of energy within the regular domain for various values ofβo (homogeneous case, horizontal excitation). . . . . . . . . . . . . . 55

2.14 Total decay of energy within the regular domain for βo = 866 s 1

(homogeneous case). . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

2.15 PML-truncated semi-infinite heterogeneous media. . . . . . . . . . . . 57

2.16 Snapshots of total displacement taken at t = 0.111 s, 0.225 s. . . . . . 59

xiii

2.17 Snapshots of total displacement taken at t = 0.233 s, and 0.290 s, onthe z = −20 m domain cross-section. . . . . . . . . . . . . . . . . . . 59

2.18 Comparison of displacement time histories between the enlarged andPML-truncated domain solutions at the sp3, sp5, sp6, and sp7 sam-pling points (heterogeneous case). . . . . . . . . . . . . . . . . . . . . 60

2.19 Relative error time history at various sampling points (heterogeneouscase). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

2.20 Total decay of energy within the regular domain for various values ofβo (heterogeneous case). . . . . . . . . . . . . . . . . . . . . . . . . . 62

2.21 Total decay of energy within the regular domain for βo = 500 s 1

(heterogeneous case). . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

2.22 Quarter model of a PML-truncated semi-infinite homogeneous media. 64

2.23 Surface load time history considered in Section 2.7.3 and its Fourierspectrum. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

2.24 Comparison of displacement time histories for various cases consideredin Table 2.4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

3.1 Time history of the Gaussian pulses and their Fourier spectrum. . . . 86

3.2 Problem configuration for the verification of the gradients. . . . . . . 89

3.3 Problem configuration: material profile reconstruction of a smoothlyvarying medium. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

3.4 Smoothly varying medium: (a) target λ and µ (MPa); and (b) profileat (x, y) = (0, 0). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

3.5 Single-parameter inversion (µ only) for a smoothly varying medium. . 93

3.6 Single-parameter inversion (λ only) for a smoothly varying medium. . 94

3.7 Variation of the misfit functional with respect to inversion iterations(single parameter inversion). . . . . . . . . . . . . . . . . . . . . . . . 95

3.8 Simultaneous inversion for λ and µ using unbiased search directions. . 96

3.9 Cross-sectional profiles of λ and µ using unbiased search directions. . 97

3.10 Simultaneous inversion for λ and µ using biased search directions. . . 98

3.11 Cross-sectional profiles of λ and µ using biased search directions. . . . 99

3.12 Variation of the misfit functional with respect to inversion iterations(simultaneous inversion). . . . . . . . . . . . . . . . . . . . . . . . . . 99

3.13 Cross-sectional profiles of cp using biased search directions. . . . . . . 100

3.14 Layered medium: (a) target λ and µ (MPa); and (b) profile at (x, y) =(0, 0). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

xiv

3.15 Simultaneous inversion for λ and µ (layered medium). . . . . . . . . . 102

3.16 Cross-sectional profiles of λ and µ (layered medium). . . . . . . . . . 103

3.17 Cross-sectional profiles of cp (layered medium). . . . . . . . . . . . . . 103

3.18 Variation of the misfit functional with respect to inversion iterations(layered medium). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

3.19 Layered medium with inclusion: (a) target λ and µ; and (b) profile at(x, y) = (7.5, 0). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

3.20 Simultaneous inversion for λ and µ (layered medium with inclusion). . 107

3.21 Cross-sectional profiles of λ and µ (layered medium with inclusion). . 108

3.22 Cross-sectional profiles of cp (layered medium with inclusion). . . . . 108

3.23 Variation of the misfit functional with respect to inversion iterations(layered medium with inclusion). . . . . . . . . . . . . . . . . . . . . 109

3.24 Measured displacement response of the layered medium with inclusion,at (x, y, z) = (3.125,−13.75, 0) m, due to the Gaussian pulse p20,contaminated with Gaussian noise. . . . . . . . . . . . . . . . . . . . 110

3.25 Simultaneous inversion for λ and µ using measured data contaminatedwith 1% and 5% Gaussian noise (layered medium with inclusion). . . 111

3.26 Simultaneous inversion for λ and µ using measured data contaminatedwith 10% and 20% Gaussian noise (layered medium with inclusion). . 112

3.27 Cross-sectional profiles of λ and µ at different noise levels (layeredmedium with inclusion). . . . . . . . . . . . . . . . . . . . . . . . . . 113

3.28 Layered medium with three inclusions: target λ and µ (a) along across-section that cuts through the domain from (x, y) = ( 20, 46.5)to ( 20, 20) to (46.5, 20); (b) along a cross-section that cuts throughthe medium from (x, y) = (20, 46.5) to (20, 20) to ( 46.5, 20); (c)profile at (x, y) = ( 20, 20); (d) profile at (x, y) = (20, 20); and (e)profile at (x, y) = (20, 20). . . . . . . . . . . . . . . . . . . . . . . . . 115

3.29 Layered medium with three inclusions: target λ and µ on (a) thez = −8.75 m cross-section; and (b) the z = −35 m cross-section. . . . 116

3.30 Simultaneous inversion for λ and µ: cross-section cuts through thedomain from (x, y) = ( 20, 46.5) to ( 20, 20) to (46.5, 20) (layeredmedium with three inclusions). . . . . . . . . . . . . . . . . . . . . . 117

3.31 Simultaneous inversion for λ and µ: cross-section cuts through thedomain from (x, y) = (20, 46.5) to (20, 20) to ( 46.5, 20) (layeredmedium with three inclusions). . . . . . . . . . . . . . . . . . . . . . 118

3.32 Layered medium with three inclusions: (a) inverted λ profile on thez = 8.75 m cross-section; (b) inverted µ profile on the z = 8.75 mcross-section; (c) inverted λ profile on the z = 35 m cross-section;and (d) inverted µ profile on the z = 35 m cross-section. . . . . . . . 119

xv

3.33 Cross-sectional profiles of λ and µ (layered medium with three inclu-sions). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

4.1 Problem definition: (a) interrogation of a heterogeneous semi-infinitedomain by an active source; (b) a 2D cross-section of the domain show-ing one source and multiple receivers; and (c) computational modeltruncated from the semi-infinite medium via the introduction of PMLs.125

4.2 A PML-truncated semi-infinite domain in two dimensions. . . . . . . 126

4.3 Approximation of a 3D halfspace problem by a 2D halfplane problem. 137

4.4 Line load with a finite length. . . . . . . . . . . . . . . . . . . . . . . 140

4.5 2D and 3D Green’s functions. . . . . . . . . . . . . . . . . . . . . . . 142

4.6 Comparison of 2D and 3D systems due to a suddenly applied load. . . 143

4.7 Comparison of 2D and 3D systems due to a rectangular pulse load. . 144

4.8 NEES@UTexas Liquidator Vibroseis. . . . . . . . . . . . . . . . . . . 145

4.9 Chirp with dominant frequencies between 3 Hz and 8 Hz. . . . . . . . 147

4.10 Line load truncation effect for different chirps. . . . . . . . . . . . . . 150

4.11 Comparison of infinite line load (L = ∞), a continuous line load offinite length (L = 100 m), and a series of point loads spaced s metersapart over a distance of 100 m. . . . . . . . . . . . . . . . . . . . . . 151

4.12 The field experiment layout. . . . . . . . . . . . . . . . . . . . . . . . 152

4.13 Hornsby Bend field experiment: (a) instrumentation van; and (b) T-rex at the site. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

4.14 Force (chirp C-3-8) applied by Liquidator at (ξ1, ξ2) = (0, 0). . . . . . 155

4.15 Velocity (due to force C-3-8 at (0, 0, 0)) measured at (−5, 0, 0). . . . . 156

4.16 Velocity (due to force C-3-8 at (0,−5, 0)) measured at (−5, 0, 0). . . . 157

4.17 Velocity (due to force C-3-8 at (0,−10, 0)) measured at (−5, 0, 0). . . 158

4.18 Equivalent line load (chirp C-3-8) applied at ξ1 = 0 m, and corre-sponding response at x1 = −5 m. . . . . . . . . . . . . . . . . . . . . 158

4.19 Inverted profiles for cp and cs at iteration 1900. . . . . . . . . . . . . 161

4.20 Inverted profile for cs via the SASW method. . . . . . . . . . . . . . . 163

4.21 Shear wave velocity profiles obtained via SASW and full-waveform-based inversion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

4.22 Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion. . . 165

4.23 Juxtaposition of CPT results and the inverted profiles. . . . . . . . . 167

xvi

4.24 The field experiment layout. . . . . . . . . . . . . . . . . . . . . . . . 168

4.25 Garner Valley field experiment: (a) T-rex at the site; (b) a buriedgeophone; (c) instrumentation van; and (d) the site. . . . . . . . . . . 169

4.26 Location of array 1 and array 2 sensors. . . . . . . . . . . . . . . . . . 171

4.27 Time-history of vertical displacement at sensors in array 1. . . . . . . 171

4.28 Time-history of vertical displacement at sensors in array 2. . . . . . . 172

4.29 Inverted profiles for cp and cs at iteration 4550 (part of the domain isnot shown to aid material visualization in depth). . . . . . . . . . . . 174

4.30 Inverted profiles for cp and cs at iteration 4550 (x = 10 m). . . . . . . 175

4.31 Inverted profiles for cp and cs at iteration 4550 (x = 60 m). . . . . . . 176

4.32 Inverted profiles for cp and cs at iteration 4550 (x = 90 m). . . . . . . 177

4.33 Garner Valley experiment: (a) layout; and (b) SASW method testlocations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

4.34 Shear wave velocity profiles of the NEES site obtained via SASW andfull-waveform-based inversion (FWI). . . . . . . . . . . . . . . . . . . 179

4.35 Inverted profile for cs via the SASW method. . . . . . . . . . . . . . . 180

4.36 Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion(x = +10m). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

4.37 Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion(x = +60m). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

4.38 Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion(x = +90m). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

4.39 Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion(x = +100m). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

xvii

Chapter 1

Introduction

A high-fidelity image of the near-surface deposits has practical significance in

the safe design of critical infrastructure components, such as bridges, hospitals, nu-

clear power plants, etc. In current practice, both invasive and non-invasive techniques

are used. Among the non-invasive techniques, wave-based approaches, whether via

elastic, acoustic, or electromagnetic waves, are dominant. Currently, the majority

of the wave-based techniques rely on simplifying assumptions that limit their appli-

cability. For instance, the widely-used Spectral-Analysis-of-Surface-Waves (SASW)

method, relies on a one-dimensional assumption that result in a horizontally layered

profile for the soil. Similarly, the Multichannel-Analysis-of-Surface-Waves (MASW)

method, though more versatile than the SASW method, falls short of reconstructing

heterogeneous three-dimensional images. By contrast, full-waveform inversion tech-

niques are inherently three- (or two-) dimensional, and can accommodate arbitrary

heterogeneity automatically.

This dissertation presents a systematic framework for the high-fidelity imag-

ing of the soil, in the context of geotechnical site characterization. The goal is to

image the arbitrarily heterogeneous material profile of a probed site, using complete

waveforms of its response to interrogating elastic waves, originating from the ground

1

surface. To this end, the response of the soil medium to active sources is collected

by receivers dispersed over the formation’s surface. Arriving at a material profile

is then accomplished by minimizing the misfit between the collected response at re-

ceiver locations, and a computed response corresponding to a trial distribution of

the material parameters. The misfit minimization problem is constrained by the

wave physics of the forward problem, resulting in a PDE-constrained optimization

problem.

Imaging near-surface deposits brings additional difficulties, typically not en-

countered in exploration geophysics, or medical imaging. Since in geotechnical site

characterization, one, typically, deals with a semi-infinite, (relatively) small domain,

an accurate domain termination tool seems necessary, and may play a critical role

in obtaining a reliable material profile. In this vein, and in the presence of hetero-

geneity, using Perfectly-Matched-Layers (PMLs) for domain termination is the best

available option, and is thus used in this work.

1.1 Background

The robust, high-fidelity subsurface imaging of the soil relies on two key com-

ponents: a) the forward problem, where a wave simulation tool for semi-infinite

domains is needed1; and b) the inverse problem, where a full-waveform inversion

approach is used for the misfit minimization. The solution of the inverse problem,

typically, necessitates an iterative process, which requires repeated solutions of the

1We use PMLs for domain truncation.

2

forward problem, thereby accentuating the importance of an efficient and accurate

forward simulator. We review next key developments in both the forward and inverse

problems, in order to place the present work in context.

1.1.1 The perfectly-matched-layer (PML)

Numerical simulation of elastic waves in unbounded heterogeneous media has

important applications in various fields, such as seismology [1], soil-structure inter-

action [2, 3], seismic imaging [4], wave-based enhanced oil recovery [5–7], and site

characterization [8]. To keep the computation feasible, one needs to limit the extent

of the computational domain. This entails considering appropriate conditions at the

truncation boundaries such that, under ideal conditions, the boundaries become in-

visible to the outgoing waves. Perfectly-matched-layers (PML) appear to be among

the best choices for domain truncation owing, especially, to their ability to handle

heterogeneity. From a practical standpoint, implementing PML in existing codes is

also easier than competitive alternatives [9, 10]. The PML is a buffer zone that en-

forces attenuation of propagating and evanescent waves. The PML’s properties vary

gradually, from a perfectly matched interface, through a progressively attenuative

medium, to, usually, a fixed termination at the buffer zone’s end2.

The PML was first introduced by Berenger for electromagnetic waves [13].

Later, it was interpreted as a mapping of the physical coordinates onto the complex

space, referred to as complex coordinate stretching [14–16]. The interpretation al-

2Other termination conditions are also possible, including local non-reflecting boundary condi-tions [11, 12].

3

lowed the further development and adoption of the PML in elastodynamics [17, 18],

for the linearized Euler equations [19], for Helmholtz equations [9], in poroelasticity

[20], and elsewhere.

Berenger’s original development, and many other early formulations, were

based on field-splitting, which partitions a physical variable into components parallel

and perpendicular to the truncation boundary; this technique alters the structure of

the underlying differential equations and results in a manyfold increase of the number

of unknowns. Gedney [21] proposed an unsplit formulation for electromagnetic waves,

citing preservation of the Maxwellian structure, and computational efficiency among

the main advantages. Abarbanel and Gottlieb showed that Berenger’s split-form is

only weakly well-posed3, and therefore is prone to instability [23]. This motivated the

development of strongly well-posed unsplit formulations [24]; however, it turned out

that the dynamical system associated with the unsplit form suffers from degeneracy

at quiescent state, which renders the scheme unstable, and further manipulation of

the equations is necessary to ensure stability [25].

In elastodynamics, Duru and Kreiss [26] proposed a well-posed discretely-

stable unsplit formulation, and mentioned that the first-order split-form is only

weakly hyperbolic4 [27]. Among other unsplit formulations, we refer to [28–30] where

the authors’ motivation stemmed primarily from exploring alternative forms, rather

than address stability. All these developments used finite differences for spatial dis-

cretization, and exploited explicit time-stepping. Among unsplit-field finite element

3See [22] for definition of well-posedness, and hyperbolicity.4Strong hyperbolicity is a desirable property and guarantees well-posedness.

4

developments, Basu and Chopra [2] presented an, almost, displacement-only proce-

dure that relies on stress-histories and needs the evaluation of an internal force vector

at every time step, as is typically done in plasticity, via an implicit time-marching

scheme based on unsymmetric matrices. Later, Basu [31] extended this work to three-

dimensional problems, using mass-lumping and explicit time-stepping. Martin et al.

[32] developed a computationally efficient procedure that couples a velocity-stress

convolutional PML (CPML) in an ad hoc manner with a displacement-only formu-

lation in the interior domain for two-dimensional problems. The CPML formulation

was used to circumvent instabilities observed when waves travel along the inter-

face between the PML and the interior domain, when the standard PML stretching

function is used. Recently, Kucukcoban and Kallivokas [33] developed a symmet-

ric displacement-stress formulation, using mixed-field finite elements for the PML,

coupled with standard displacement-only finite elements for the interior domain, us-

ing the standard Newmark method for time integration. We remark that implicit

time-stepping can become challenging for large-scale three-dimensional problems and

should be avoided if possible.

The literature on split-field elastodynamics is rich. This approach is particu-

larly attractive because, normally, it does not use convolutions or auxiliary variables.

However, it almost always results in using mixed schemes, i.e., treating velocity and

stress components (or a similar combination) as unknowns over the entire domain.

Table 1.1 summarizes key developments in time-domain elastodynamics based on four

categories: split- or unsplit-field formulation, and finite difference or finite/spectral

element implementation.

5

Table 1.1: PML developments in time-domain elastodynamics.

Split-field Unsplit-field

FD Chew and Liu[17] Wang and Tang [30]Hastings et al. [34] Drossaert and Giannopoulos [28]Liu [35] Komatitsch and Martin [29]Collino and Tsogka [18] Duru and Kreiss [26]

FE/SE Collino and Tsogka [18] Basu and Chopra [2]Becache et al. [36] Martin et al. [32]Festa and Nielsen [37] Basu [31]Komatitsch and Tromp [38] Kucukcoban and Kallivokas [33]Cohen and Fauqueux [39]Festa and Vilotte [40]Meza-Fajardo and Papageorgiou [41]

Differences between various PML formulations are not only due to the split

or unsplit formulation and numerical implementation, but also on the choice of co-

ordinate stretching function. The classical stretching function has been criticized for

allowing spurious growths in numerical simulations in two dimensions, when waves

impinge at grazing incidence on the PML interface. These growths have been loosely

attributed to a zero-frequency singularity in the classical stretching function, and

have been, reportedly, alleviated by using a complex-frequency-shifted (CFS) stretch-

ing function, which removes the singularity [28]. However the CFS-PML loses its

absorptive competence at low frequencies [42]. Meza-Fajardo and Papageorgiou [41]

proposed a multi-axial stretching approach and demonstrated its successful perfor-

mance for waves impinging on the PML at grazing incidence, as well as for problems

involving anisotropy. In [43, 44], it was reported that the M-PML is not perfectly-

matched at the interface; however, later on, Meza-Fajardo and Papageorgiou showed

that the multi-axial perfectly-matched-layer (M-PML) is indeed perfectly-matched

6

in Berenger’s sense, and it provides domain truncations that are at least as accu-

rate as the classical PML, when the latter is stable [45]. In a more recent study,

Ping et al. [46] have shown results according to which M-PML may perform less

accurately than the classical PML. Our own experience, both in two-dimensional

and three-dimensional simulations, is also more in accordance with Ping et al. [46].

It also seems that the original M-PML development is mathematically inconsistent

due to the improper definition of the Jacobian of the transformation. We discuss

this issue in Section 2.6. Herein, we opt for classical stretching functions for their

simplicity, satisfactory performance when parametrized carefully, and their accuracy

in low frequencies, which is important in site characterization problems [8]. We also

discuss how our formulation can accommodate the multi-axial stretching through

simple modifications.

1.1.2 Full-waveform inversion (FWI)

Seismic inversion refers to the process of identification of material properties in

geological formations [47–49]. The problem arises predominantly in exploration geo-

physics [50–53] and geotechnical site characterization [8]; it belongs to the broader

class of inverse medium problems: waves, whether of acoustic, elastic, or electro-

magnetic nature, are used to interrogate a medium, and the medium’s response

to the probing is subsequently used to image the spatial distribution of properties

(e.g., Lame parameters, or wave velocities) [54–56]. Mathematically, algorithmically,

and computationally, inverse medium problems are challenging, especially, when no

a priori constraining assumption is made on the spatial variability of the medium’s

7

properties. The challenges are further compounded when the underlying physics is

time-dependent, and involves more than a single distributed parameter to be inverted

for, as in seismic inversion.

Due to the complexity of the inverse problem at hand, most techniques to date

rely on simplifying assumptions, aiming at rendering a solution to the problem more

tractable. These assumptions can be divided into four categories: a) assumptions

regarding the dimensionality of the problem, whereby the original problem is reduced

to a two-dimensional [8, 57], or a one-dimensional problem [58]; b) assuming that the

dominant portion of the wave energy on the ground surface is transported through

Rayleigh waves, and thus, disregarding other wave types, such as compressional

and shear waves, as is the case in the Spectral-Analysis-of-Surface-Waves (SASW)

and its variants (MASW) [59]; c) inverting for only one parameter, as is done in

[60–63], where inversion was attempted only for the shear wave velocity, assuming

the compressional wave velocity (or an equivalent counterpart) is known; and d)

assumptions concerning the truncation boundaries, which are oftentimes, grossly

simplified due to the complexity associated with the rigorous treatment of these

boundaries [64]. Over the past decade, continued advances in both algorithms and

computer architectures have allowed the gradual removal of the limitations of existing

methodologies. However, a robust methodology, especially for the time-dependent

elastic case remains, by and large, elusive.

Among the recent works on inversion, which are similar in character to ours,

we refer to Pratt et al. [65] who considered two-dimensional acoustic inversion in

the frequency domain, and Epanomeritakis et al. [61] where full-waveform inversion

8

has attempted for three-dimensional time-domain elastodynamics, where a simple

boundary condition was used for domain truncation. Kang and Kallivokas [56] con-

sidered the problem for the two-dimensional time-domain acoustic case, and used

PMLs to accurately account for domain truncation. Kucukcoban [57] extended the

work of Kang and Kallivokas to two-dimensional elastodynamics, and reported suc-

cessful reconstruction of the two Lame parameters for models involving synthetic

data. Recently, Bramwell [66] used a discontinuous Petrov-Galerkin (DPG) method

in the frequency domain, endowed with PMLs, for seismic tomography problems,

advocating the DPG scheme over conventional continuous Galerkin methods, since

it results in less numerical pollution.

This dissertation extends the work of Kucukcoban [57] to three-dimensional

elastodynamics. We remark that PMLs add significant complexity to the solution

of both the forward and the inverse problem. Moreover, since we target three-

dimensional problems, using scalable parallel algorithms is essential.

1.2 Present approach

In this dissertation, we discuss a systematic framework for the numerical res-

olution of the inverse medium problem, directly in the time-domain, in the context

of geotechnical site characterization. As discussed in the introduction, the goal is to

image the arbitrarily heterogeneous material profile of a probed soil medium, using

complete waveforms5 of its response to interrogating elastic waves, originating from

5Using the complete waveform (complete recorded response) results in a full-waveform inversionapproach.

9

the ground surface. To this end, the response of the soil medium to active sources

(Vibroseis equipment) is collected by receivers (geophones) dispersed over the for-

mation’s surface, as shown in Figure 1.1(a). Arriving at a material profile is then

accomplished by minimizing the difference between the collected response at receiver

locations, and a computed response corresponding to a trial distribution of the ma-

terial parameters. Due to the heterogeneity, we use PMLs for domain termination,

as the best available option. Figure 1.1 shows the prototype computational model,

where the physical domain has been replaced by a computational domain terminated

by PMLs at the truncation boundaries.

(a)

x y

z

(b)

Figure 1.1: Problem definition: (a) interrogation of a heterogeneous semi-infinitedomain by an active source; and (b) computational model truncated from the semi-infinite medium via the introduction of PMLs.

In order to address all the difficulties outlined earlier, we integrate recent ad-

vances in several areas. Specifically, we use (a) a parallel, state-of-the-art wave sim-

ulation tool for domains terminated by PMLs [67]; (b) a partial-differential-equation

10

(PDE)-constrained optimization framework through which the minimization of the

difference between the collected response at receiver locations and a computed re-

sponse corresponding to a trial distribution of the material properties is attained [68];

(c) regularization schemes to alleviate the ill-posedness inherent in inverse problems;

(d) continuation schemes that lend algorithmic robustness [56]; and (e) a biasing

scheme that accelerates the convergence of the λ-profile for robust simultaneous in-

version of both Lame parameters [57].

1.3 Contributions

This work builds and improves upon the Ph.D. dissertation of Sezgin Kucuk-

coban in two-dimensional elastic full-waveform inversion [57]. Key contributions of

the present development are listed below.

In three dimensions:

• Developing a new PML formulation for the simulation of elastic waves in three-

dimensional, arbitrarily heterogeneous, semi-infinite media, where a stress-

displacement formulation for the PML, is coupled with a standard displacement-

only formulation for the interior domain. This hybrid treatment leads to a

computationally cost-efficient scheme. The formulation builds and improves

upon a recently developed two-dimensional scheme [33]. However, it is restruc-

tured, and modified, to accommodate explicit time-stepping, which makes it

suitable for large-scale problems on parallel computers [67].

• Presenting a robust full-waveform inversion methodology for three-dimensional,

11

arbitrarily heterogeneous, PML-truncated, elastic formations, leading to the

successful reconstruction of the spatially-distributed Lame parameters. A con-

sistent finite element approach was used throughout. The accuracy of the

discrete gradients, computed from this scheme, are verified by comparing them

with directional finite differences [69]. The developed framework was used

for the three-dimensional characterization of the NEES@UCSB site in Garner

Valley, CA [70].

In two dimensions:

• Developing a discretize-then-optimize scheme for the accurate computation of

the discrete gradients of the discrete objective functional. Accordingly, the

objective functional is discretized first, followed by differentiation, to yield

discrete gradients, which can then be used in a gradient-based optimization

scheme [8].

• Developing a practical procedure to accommodate field data, which are inher-

ently three-dimensional, into two-dimensional full-waveform-inversion-based co-

des. Designing and conducting a field experiment at the Hornsby Bend site in

Austin, TX, whose records were subsequently used to drive the inversion al-

gorithms in order to characterize the site where the experiment took place

[8].

1.4 Dissertation outline

The rest of this dissertation is organized as follows:

12

Chapter 2 presents a new PML formulation for the simulation of elastic waves

in three-dimensional, arbitrarily heterogeneous domains. We begin with reviewing

key ideas for developing a PML, and discuss complex coordinate stretching. Then, we

present a stress-displacement formulation for the PML, which leads to a third-order-

in-time semi-discrete form. We discuss how this formulation can be coupled with

a standard displacement-only formulation for the interior domain, thus leading to

a computationally cost-efficient scheme. We discuss several time-marching schemes.

In particular, we discuss an explicit, fourth-order, Runge-Kutta scheme which is

well-suited for large-scale problems on parallel computers. In Section 2.5, we discuss

an alternative formulation for the PML that leads to a symmetric semi-discrete

form. In Section 2.6 we show how our formulation can accommodate multi-axial

PML (M-PML), by simple modifications. Lastly, we provide numerical experiments

demonstrating stability and efficacy of the proposed formulations.

In Chapter 3, we consider the inverse medium problem in three-dimensional,

PML-truncated domains, using full-waveforms. We cast the associated inverse prob-

lem, as a misfit minimization problem, using the apparatus of PDE-constrained opti-

mization to impose the forward wave propagation equations, followed by computing

the optimality system. Next, we discuss strategies that alleviate ill-posedness, and

lend algorithmic robustness to our proposed inversion scheme. By using a numerical

experiment, we verify the accuracy of the gradients computed via the control prob-

lems, by comparing them with directional finite differences. We present numerical

experiments demonstrating successful reconstruction of the two Lame parameters for

smooth and sharp profiles, using noise-free and also highly-noisy synthetic data.

13

In Chapter 4, we discuss a full-waveform inversion methodology for site char-

acterization, using field data. We start by reviewing the two-dimensional forward

wave propagation in PML-truncated domains, followed by presenting a robust ap-

proach to tackle the associated inverse medium problem. We then report on the

design and data processing of a field experiment, whose records were used along

with the presented two-dimensional framework, to obtain the compressional, and

shear wave velocity profile of the site where the experiment took place. Next, we

compare the profiles with those obtained from the SASW method, and invasive Cone

Penetrometer Tests (CPTs). Lastly, we use the methodology described in Chapter

3 for the three-dimensional site characterization of the NEES site in Garner Valley,

CA.

We conclude with summary remarks in Chapter 5, and suggest future direc-

tions.

14

Chapter 2

Simulation of wave motion in three-dimensional

PML-truncated heterogeneous media

In this chapter, we discuss the development and parallel implementation of an

unsplit-field, displacement-stress PML formulation, using mixed-field finite elements

for the PML, which when coupled with a standard displacement-only finite element

formulation for the interior domain, leads to the efficient simulation of wave motion

in physically unbounded, three-dimensional, arbitrarily heterogeneous elastic media.

The hybrid treatment of coupling a mixed-field PML with a single-field interior-

domain leads to optimal computational cost and allows for ready incorporation of

the PML in existing closed-domain standard finite element codes, by simply attaching

the matrices corresponding to the PML buffer. The resulting semi-discrete form is

unsymmetric and third-order in time. Using spectral elements, we render the mass

matrix diagonal and exploit explicit time-stepping via the Runge-Kutta method. We

also present an alternative formulation, which results in a fully symmetric discrete

form, at the expense of utilizing an implicit time-marching scheme. We discuss how

the standard Newmark scheme can also be used for time integration. This work builds

and improves upon recent developments [33, 71] in two-dimensional elastodynamics.

15

2.1 Complex-coordinate-stretching

In this section, we briefly review the key features of the PML. Part of the

material discussed here is not new; however, it is provided to allow for context and

completeness.

2.1.1 Key idea

The key idea in constructing a PML is based on analytic continuation of so-

lutions of wave equations. This amounts to mapping the spatial coordinates onto

the complex space, using the, so-called, stretching functions. For instance, one-

dimensional outgoing waves propagate according to uout(x, t) = e−ik(x−ct), where

k is the wavenumber, and c denotes wave speed. After applying the mapping1

x 7→ ζ(x) + 1iω

η(x), we obtain uoutPML(x, t) = e−ik(ζ(x)−ct)e−η(x)/c, where the lat-

ter term enforces spatial attenuation. A similar argument also holds for evanescent

waves.

In practice, the PML has a limited thickness (see Fig. 2.1), and is termi-

nated with a fixed boundary. Therefore, reflections (i.e., incoming waves) could

develop when outgoing waves hit the fixed boundary of the PML layer. In our

one-dimensional example, uincPML(x, t) = eik(ζ(x)+ct)eη(x)/c. Since η(x) is a positive,

monotonically increasing function of x, reflected waves also get attenuated, due to

decreasing x. Hence, the PML attenuates both outgoing and incoming waves.

We briefly discuss the principal components required for constructing a PML.

1ζ(x), η(x) are positive, monotonically increasing functions of x.

16

Referring to Fig. 2.1, let s denote the coordinate variable normal to the interface

of the interior domain with the PML. For the interior domain it holds 0 < s < s0,

whereas for the PML s0 < s < st; LPML denotes the thickness of the PML layer,

and ns is the outward unit normal at the interface, pointing away from the interior

domain. The physical coordinate s is mapped (or “stretched”) to s within the PML

region according to

s 7→ s = so +

∫ s

so

λs(s′, ω) ds′, (2.1)

where ω denotes circular frequency, and λs is the, so-called, stretching funcion.

Interiordomain

PML

Outgoing wave

Attenuated

Reflected

LPMLL

s

ns

sost

0

Figure 2.1: A PML truncation boundary in the direction of coordinate s.

The classical PML results from choosing the stretching function according to

λs(s, ω) = αs(s) +1

iωβs(s), (2.2)

17

where αs is the scaling function, which stretches the coordinate variable s, whereas

βs is the attenuation function, which enforces the amplitude decay of propagating

waves2. For evanescent waves, αs improves amplitude decay by elongating the real

coordinate variable s. For the interface to be “invisible” to the waves entering the

PML (perfect matching), αs|s=s0 = 1, and βs|s=s0 = 0. Moreover, αs and βs are

positive, non-decreasing functions of s. Finally, applying the fundamental theorem

of calculus to (2.1), there results

ds

ds=

d

ds

∫ s

so

λs(s′, ω) ds′ = λs(s, ω). (2.3)

Hence, we obtain the following derivative rule between the stretched coordinate sys-

tem, and the physical coordinate system

d(·)ds

=1

λs(s, ω)

d(·)ds

. (2.4)

The PML governing equations are naturally written in the stretched coordi-

nate system. We frequently use (2.4) to express the PML equations in the physical

coordinate system.

2.1.2 Choice of stretching functions

The main requirements for the stretching functions are a) perfect matching at

the interface; b) positive non-decreasing variability; and c) a gradual and smoothly

2In our one-dimensional example, we used the notation ζ(x) =∫ x

0α(x′)dx′, and

η(x) =∫ x

0β (x′) dx′.

18

varying profile. The last requirement is particularly important for numerical dis-

cretization, since, for adequate resolution, a sharply varying profile requires a finer

mesh than a smoother profile. A widely adopted form that satisfies these require-

ments is given in terms of polynomials, as in

αs(s) = 1 + αo

[(s− so)ns

LPML

]m, so ≤ s ≤ st, (2.5a)

βs(s) = βo

[(s− so)ns

LPML

]m, so ≤ s ≤ st, (2.5b)

where α0 and β0 are user-tunable parameters that control amplitude decay, and m

denotes polynomial degree.

For one-dimensional problems, prior to discretization, β0 can be shown to be

β0 =(m+ 1) cp2 LPML

log

(1

R

), (2.6)

where R is the amount of reflection from the fixed PML boundary, and cp is the

P-wave velocity. In practice, however, selecting appropriate values for α0 and β0 is

not straightforward. The choice depends on the problem at hand, mesh resolution,

and, it typically, needs a few experiments to be optimized. We remark that the

performance of the PML relies heavily on its careful parameterization [11, 33].

We also worked with the following trigonometric profiles that are smoother

than polynomials; however, we did not observe any compelling improvement.

19

αs(s) = 1 +αo

2

[1 + sin

(π( |s− so|

LPML− 1

2

))], so ≤ s ≤ st, (2.7a)

βs(s) =βo

2

[1 + sin

(π( |s− so|

LPML

− 1

2

))], so ≤ s ≤ st. (2.7b)

2.2 Three-dimensional unsplit-field PML

The linear elastic wave equation, in the absence of body forces, can be written

as the following system

div ST = ρu, (2.8a)

S = µ[∇u+ (∇u)T

]+ λ(divu)I, (2.8b)

where (2.8a) represents conservation of linear momentum, and (2.8b) is the combined

constitutive and kinematic equations; S represents the Cauchy stress tensor, u is the

displacement vector, ρ denotes mass density of the medium, λ and µ are the two Lame

parameters, I is the second-order identity tensor, and a dot (˙) denotes differentiation

with respect to time of the subtended variable.

To derive the corresponding PML equations, we first Fourier-transform (2.8)

with respect to the time variable. Writing the resulting differential equations in

the stretched coordinate system affords the sought-after spatial decaying property.

These equations can then be expressed in the physical coordinate system by using

(2.4). Finally, exploiting the inverse Fourier transform results in the corresponding

time-domain equations.

20

2.2.1 Frequency-domain equations

We Fourier-transform (2.8) with respect to the time variable; there results:

div ST = (iω)2ρu, (2.9a)

S = µ[∇u+ (∇u)T

]+ λ(div u)I, (2.9b)

where a caret (ˆ) denotes the Fourier transform of the subtended variable, and spatial

and frequency dependency of the variables are suppressed for brevity. We focus on

(2.9a) first, and express it in the stretched coordinate system by replacing x, y, and

z with x, y, and z, respectively. For clarity, we use the unabridged notation:

∂Sxx

∂x+

∂Syx

∂y+

∂Szx

∂z= (iω)2ρ ux, (2.10a)

∂Sxy

∂x+

∂Syy

∂y+

∂Szy

∂z= (iω)2ρ uy, (2.10b)

∂Sxz

∂x+

∂Syz

∂y+

∂Szz

∂z= (iω)2ρ uz, (2.10c)

where Sij , and ui, denote stress tensor, and displacement vector components, respec-

tively. Equation (2.10) can be expressed in the physical (un-stretched) coordinate

system by using (2.4); thus, we obtain:

21

1

λx

∂Sxx

∂x+

1

λy

∂Syx

∂y+

1

λz

∂Szx

∂z= (iω)2ρ ux, (2.11a)

1

λx

∂Sxy

∂x+

1

λy

∂Syy

∂y+

1

λz

∂Szy

∂z= (iω)2ρ uy, (2.11b)

1

λx

∂Sxz

∂x+

1

λy

∂Syz

∂y+

1

λz

∂Szz

∂z= (iω)2ρ uz. (2.11c)

Multiplying (2.11) through by λxλyλz results in

div(STΛ

)= (iω)2ρ λxλyλz u, (2.12)

where the stretching tensor Λ is defined as

Λ =

λyλz 0 00 λxλz 00 0 λxλy

=

αyαz 0 00 αxαz 00 0 αxαy

+1

(iω)

αyβz + αzβy 0 0

0 αxβz + αzβx 00 0 αxβy + αyβx

+1

(iω)2

βyβz 0 00 βxβz 00 0 βxβy

= Λe +

1

iωΛp +

1

(iω)2Λw. (2.13)

We remark that within the interior domain, Λe reduces to the identity ten-

sor, whereas Λp and Λw vanish identically. Substituting (2.13) and (2.2) in (2.12),

rearranging and grouping similar terms, results in

div

(STΛe +

1

iωSTΛp +

1

(iω)2STΛw

)= ρ

[(iω)2au+ iωbu+ cu+

d

iωu

], (2.14)

22

where

a = αx αy αz,

b = αx αy βz + αx αz βy + αy αz βx,

c = αx βy βz + αy βz βx + αz βy βx,

d = βx βy βz. (2.15)

Multiplying (2.14) by iω, we obtain

div

(iωSTΛe + STΛp +

1

iωSTΛw

)= ρ

[(iω)3au+ (iω)2bu+ iωcu+ du

]. (2.16)

Next, we focus our attention on the combined constitutive and kinematic

equations (2.9b). Writing (2.9b) in the stretched coordinate system, and using (2.4)

to express it in the physical coordinate system, there results

S = µ

(∇u)

1λx

0 0

0 1λy

0

0 0 1λz

+

1λx

0 0

0 1λy

0

0 0 1λz

(∇u)T

+ λ( 1

λx

∂ux

∂x+

1

λy

∂uy

∂y+

1

λz

∂uz

∂z

)I. (2.17)

Multiplying (2.17) by λxλyλz results in

λxλyλzS = µ[∇u Λ + Λ (∇u)T

]+ λ div(Λu)I, (2.18)

where the stretching tensor Λ is defined in (2.13). Multiplying (2.18) by (iω)2 and

using (2.13), and (2.2), rearranging and grouping similar terms, we obtain

23

(iω)2aS+ iωbS + cS+1

iωdS = µ(iω)2

[(∇u)Λe + Λe(∇u)T

]

+ µ iω[(∇u)Λp + Λp(∇u)T

]+ µ

[(∇u)Λw + Λw(∇u)T

]

+ λ(iω)2 div(Λeu)I+ λ iω div(Λpu)I+ λ div(Λwu)I. (2.19)

Equations (2.16) and (2.19) constitute the corresponding frequency-domain momen-

tum, and combined constitutive and kinematic equations in the stretched coordinate

system, respectively. They possess the desired spatial decaying property.

2.2.2 Time-domain equations

In this section, we apply the inverse Fourier transform to (2.16) and (2.19)

to obtain the corresponding time-domain equations. This operation is rather simple

due to the specific choice of the stretching function (2.2). We use

F−1

[g(ω)

]=

∫ t

0

g(τ)dτ, (2.20)

where F−1 denotes the inverse Fourier transform operator3, and g(t) is a sufficiently

regular function. Applying the inverse Fourier transform to (2.16) and (2.19), we

obtain

3In general, F−1[g(ω)iω

]=

∫ t

0 g(τ)dτ−πg(0)δ(ω), but, it can be shown that since, by construction,

the overall development excludes ω = 0, the inverse transform reduces to (2.20) [71].

24

div

[STΛe + STΛp +

(∫ t

0

STdτ

)Λw

]= ρ (a

...u+ bu+ cu+ du) , (2.21a)

aS+ bS + cS+ d

(∫ t

0

Sdτ

)=

µ[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]+

λ [div(Λeu) + div(Λpu) + div(Λwu)] I. (2.21b)

The set of integro-differential equations (2.21) can be expressed as a set of

only partial differential equations, upon introducing the auxiliary variable S(x, t),

which may be interpreted as the stress history tensor [71, 72]:

S(x, t) =

∫ t

0

S(x, τ)dτ. (2.22a)

Clearly,

S(x, t) = S(x, t), S(x, t) = S(x, t),...S(x, t) = S(x, t). (2.22b)

Substituting (2.22) in (2.21), we obtain

div(STΛe + STΛp + STΛw

)= ρ (a

...u+ bu+ cu+ du) , (2.23a)

a...S+ bS+ cS+ dS =

µ[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]+

λ [div(Λeu) + div(Λpu) + div(Λwu)] I. (2.23b)

25

Equations (2.23) constitute the corresponding time-domain PML momentum,

and combined constitutive and kinematic equations.

2.3 Hybrid finite element implementation

In this section, we discuss an efficient finite element technique for transient

elastodynamics in PML-truncated domains. We use a method-of-lines approach,

where we exploit a Galerkin method for spatial discretization, thus, obtaining a

third-order, continuous-in-time system of ordinary differential equations. Various

methods exist for time-integration of such systems. We discuss three techniques that

seem suitable for practical applications.

2.3.1 Spatial discretization

The PML equations (2.23) can be used both for the interior domain and the

PML buffer zone, since by construction, they reduce to (2.8) in the interior domain.

This unified treatment amounts to considering stress and displacement components

as unknowns in both the interior domain and the PML buffer zone. While feasible

in principle, as is done in most PML formulations to date, we opt for a hybrid treat-

ment, originally developed in [33] for two-dimensional problems, where the interior

domain is treated with a standard displacement-only formulation, coupled with the

PML equations in the buffer zone. This approach results in substantial reduction

in computational cost compared to mixed-field formulations cast over the entire do-

main. It also makes the modification of existing interior-domain elastodynamic codes

straightforward, since, one needs to only add the PML-related forms, whereas for the

26

most part, the general structure of such codes remains intact.

Accordingly, find u(x, t) in ΩRD ∪ΩPML, and S(x, t) in ΩPML (see Fig. 2.2 for

domain and boundary designations), where u and S reside in appropriate function

spaces, and:

divµ[∇u+ (∇u)T

]+ λ(div u)I

+ b = ρ

...u in ΩRD × J,

(2.24a)

div(STΛe + STΛp + STΛw

)= ρ (a

...u+ bu+ cu+ du) in ΩPML × J,

(2.24b)

a...S+ bS+ cS+ dS =

µ[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]+

λ [div(Λeu) + div(Λpu) + div(Λwu)] I in ΩPML × J.(2.24c)

The system is initially at rest, and subject to the following boundary and interface

conditions:

µ[∇u+ (∇u)T

]+ λ(div u)I

n+ = gn on ΓRD

N × J,(2.25a)

(STΛe + STΛp + STΛw)n− = 0 on ΓPML

N × J,(2.25b)

u = 0 on ΓPMLD × J,(2.25c)

u+ = u− on ΓI × J,(2.25d)

µ[∇u+ (∇u)T

]+ λ(div u)I

n+ + (STΛe + STΛp + STΛw)n

− = 0 on ΓI × J,(2.25e)

27

where temporal and spatial dependencies are suppressed for brevity; ΩRD denotes the

interior (regular) domain, ΩPML represents the region occupied by the PML buffer

zone, ΓI is the interface boundary between the interior and PML domains, ΓRDN

and ΓPMLN denote the free (top surface) boundary of the interior domain and PML,

respectively, and J = (0, T ] is the time interval of interest. In (2.24a), b denotes

body force per unit volume.

n

y

z

Figure 2.2: PML-truncated semi-infinite domain.

We remark that the temporal differentiation in (2.24a) is necessary for the

boundary integrals that result from the weak form of (2.24a) and (2.24b) to cancel

out; this is manifested in (2.25e) and enforces the balance of tractions at the inter-

face of the interior domain and the PML. Moreover, (2.25d) implies continuity of

displacements at the interface; (2.25a) specifies tractions (gn) on the top surface of

the interior domain, and (2.25b) implies traction-free boundary condition on the top

PML surface. We consider fixed boundaries for the PML on the sides and at the

bottom, as indicated by (2.25c); other possibilities also exist but they seem to have

28

little influence on performance [11, 12].

Next, we seek a weak solution, corresponding to the strong form of (2.24)

and (2.25), in the Galerkin sense. Specifically, we take the inner products of (2.24a)

and (2.24b) with (vector) test function w(x), and integrate by parts over their cor-

responding domains. Incorporating (2.25d-2.25e) eliminates the interface boundary

terms and results in (2.26a). Next, we take the inner product of (2.24c) with (tensor)

test function T(x); there results (2.26b). There are other possibilities for deriving a

weak form that corresponds to the strong form (2.24) and (2.25). We refer to [73]

for further details.

Accordingly, find u ∈ H1(Ω)× J, and S ∈ L2(Ω)× J, such that:

ΩRD

∇w :µ[∇u+ (∇u)T

]+ λ(div u)I

dΩ +

ΩPML

∇w :(STΛe + STΛp + STΛw

)dΩ

+

ΩRD

w · ρ...u dΩ +

ΩPML

w · ρ (a...u+ bu+ cu+ du) dΩ =

ΓRD

N

w · gn dΓ +

ΩRD

w · b dΩ,

(2.26a)∫

ΩPML

T :(a...S+ bS+ cS+ dS

)dΩ

=

ΩPML

T :µ[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]

+T :λ [div(Λeu) + div(Λpu) + div(Λwu)] I dΩ, (2.26b)

for every w ∈ H1(Ω) and T ∈ L2(Ω), where gn ∈ L2(Ω) × J, and b ∈ L2(Ω) × J.

Function spaces for scalar- (v), vector- (v), and tensor-valued (A) functions are

defined as

29

L2(Ω) =

v :

Ω

|v|2dx <∞, (2.27a)

L2(Ω) =v : v ∈ (L2(Ω))3

, (2.27b)

L2(Ω) =A : A ∈ (L2(Ω))3×3

, (2.27c)

H1(Ω) =

v :

Ω

(|v|2 + |∇v|2

)dx <∞, v|ΓPML

D

= 0

, (2.27d)

H1(Ω) =v : v ∈ (H1(Ω))3

. (2.27e)

In order to resolve (2.26) numerically, we use standard finite-dimensional sub-

spaces. Specifically, we introduce finite-dimensional subspaces Ξh ⊂ H1(Ω) and

Υh ⊂ L2(Ω), with basis functions Φ and Ψ, respectively. We then approximate

u(x, t) with uh(x, t) ∈ Ξh × J, and S(x, t) with Sh(x, t) ∈ Υh × J, as detailed below

uh(x, t) =

ΦT (x)ux(t)ΦT (x)uy(t)ΦT (x)uz(t)

, (2.28a)

Sh(x, t) =

ΨT (x)Sxx(t) ΨT (x)Sxy(t) ΨT (x)Sxz(t)ΨT (x)Syx(t) ΨT (x)Syy(t) ΨT (x)Syz(t)ΨT (x)Szx(t) ΨT (x)Szy(t) ΨT (x)Szz(t)

. (2.28b)

In a similar fashion, we approximate the test functions, w(x) with wh(x) ∈

Ξh, and T(x) with Th(x) ∈ Υh; therefore:

30

wh(x) =

wT

xΦ(x)wT

y Φ(x)wT

z Φ(x)

, (2.29a)

Th(x) =

TT

xxΨ(x) TTxyΨ(x) TT

xzΨ(x)TT

yxΨ(x) TTyyΨ(x) TT

yzΨ(x)TT

zxΨ(x) TTzyΨ(x) TT

zzΨ(x)

. (2.29b)

Incorporating (2.28-2.29) into (2.26), results in the following semi-discrete form

M...d+Cd+Kd+Gd = f , (2.30)

where spatial and temporal dependencies are suppressed for brevity, and system

matrices, M, C, K, G, and vectors d and f , are defined as

M =

[MRD + Ma 0

0 Na

], C =

[Mb Aeu

−ATel Nb

], (2.31a)

K =

[KRD + Mc Apu

−ATpl Nc

], G =

[Md Awu

−ATwl Nd

], (2.31b)

d =[uh Sh

]T, f =

[fRD 0

]T, (2.31c)

where subscript RD refers to the interior (regular) domain, and MRD, KRD, and fRD,

correspond to the standard mass matrix, stiffness matrix, and vector of nodal forces in

the interior domain, respectively, and a bar indicates their extension to encompass

all the displacement degrees-of-freedom4; uh and Sh comprise the vector of nodal

4This is merely a formalism to arrive at a unified, yet informative, matrix representation. Forinstance, we take KRD and extend it by adding zero entries corresponding to the uh componentsof the PML buffer. This makes the matrix-vector operation KRD uh meaningful, where, now, uh

contains the displacement degrees-of-freedom of the entire domain.

31

displacements and stresses. Moreover, uh is partitioned such that its first entries

belong solely to the interior domain, followed by those on the interface boundary

between the interior domain and the PML buffer, and finally those that are located

only within the PML. The rest of the submatrices in (2.31) correspond to the PML

buffer zone (see Fig. 2.3 for a schematic partitioning, and Appendix A.1 for submatrix

definitions; the dotted line in Fig. 2.3 separates displacement from stress degrees-of-

freedom).

Figure 2.3: Partitioning of submatrices in (2.31b).

We remark that the upper-left corner blocks of M and K correspond to the

mass and stiffness matrices of a standard displacement-only formulation, as depicted

in Fig. 2.3. This implies that in order to accommodate PML capability into existing

codes, one needs to account only for the submatrices on the lower-right blocks of M,

C, K, G.

The matrix M has a block-diagonal structure (see (A.2a)-(A.2c)); thus, it can

be diagonalized if one employs spectral elements, which then enables explicit time

integration of (2.30): this is discussed in Section 2.4.

32

Notice that the semi-discrete form (2.30) is not symmetric. In fact, a block-

diagonal structure for M comes at the price of losing symmetry. Alternatively, one

may preserve symmetry of the matrices in the semi-discrete form at the expense of

losing the block-diagonal form ofM, and thus the ability for explicit time integration.

We discuss this alternative formulation in Section 2.5.

2.3.2 Discretization in time

In this section, we discuss various possibilities of integrating the semi-discrete

form (2.30) in time. One may apply a time-marching scheme directly to (2.30), which

is third-order in time, or, exploit a more common scheme by first expressing (2.30)

as a second- or first-order in time system, via the introduction of auxiliary vectors.

Time-integration can be accomplished by working with either (2.30) or one of

its second- or first-order system counterparts, or, alternatively, one may (analytically)

integrate (2.30) in time first, to obviate the temporal differentiation of the forcing

vector. Assuming the system is initially at rest, there results

Md+Cd+Kd+Gd = f , (2.32a)

d =

∫ t

0

d(τ)|PML dτ, (2.32b)

where d is the vector of history terms. Equation (2.32) can be integrated via an

extended Newmark method as outlined in Appendix B.2. The scheme is implicit and

requires matrix factorization.

33

We remark that d contains displacement and stress degrees-of-freedom that

are associated with the PML buffer only ; therefore, its size is much smaller than d

(Fig. 2.3).

Alternatively, (2.32) can be expressed as a second-order system

Md+Cd+Kd+Gd = f , (2.33a)

˙d = d|PML. (2.33b)

In matrix notation, (2.33) reads

[M 00 0

][d¨d

]+

[C 00 I

] [d˙d

]+

[K G−I 0

] [dd

]=

[f0

], (2.34)

where now a standard Newmark scheme may be utilized to integrate (2.34); or,

alternatively

[M 00 I

][d¨d

]+

[C 0−I 0

] [d˙d

]+

[K G0 0

] [dd

]=

[f0

], (2.35)

where the resulting system can be integrated explicitly, provided that M is diagonal,

as we discuss in Section 2.4.

One may also express (2.32) as a first-order system

d

dt

x0

x1

Mx2

=

0 I 00 0 I−G −K −C

x0

x1

x2

+

00f

, (2.36)

34

where x0 = d, x1 = d, and x2 = d. Various standard explicit schemes could

then be used, provided that M is diagonal [74]. Here, we favor an explicit fourth-

order Runge-Kutta (RK-4) method. Based on various numerical experiments we

performed, we found out that, for the RK-4, ∆t < 0.8∆xcp

ensures stability on uniform

grids, where ∆x is the minimum distance between two grid points, and cp is the

maximum compressional wave velocity over an element. If, for a certain choice of

time step, a simulation with displacement-only finite elements is stable, then, the

associated simulation involving the PML is also stable with the same time step. In

other words, the introduction of the PML does not impose a more onerous time step

choice than an interior elastodynamics problem would require.

2.4 Spectral elements and explicit time integration

Hyperbolic initial-value-problems are, in general, advanced in time by using

explicit methods [22, 75]. This obviates the need for “inverting” a large linear system,

typically encountered in implicit schemes. Moreover, explicit schemes naturally lend

themselves to parallel computation, which is essential when dealing with large-scale

simulations in three-dimensional problems. In this section, we discuss how the matrix

M in the semi-discrete form (2.30) may be diagonalized, thus, enabling explicit time-

stepping via the techniques discussed in Section 2.3.2.

The simplest way of obtaining (discrete) diagonal mass-like matrices, is by

mass-lumping, as was done in [31, 76] where the authors used linear elements5. To

5By contrast to classical Galerkin finite elements, a finite difference formulation automaticallyyields diagonal mass-like matrices; see [74] for instance.

35

achieve high-order accuracy, however, one may use nodal spectral elements, where

numerical integration (quadrature rule) is based on the same nodes that polyno-

mial interpolation is carried out [77, 78]. This results in (discrete) diagonal mass-like

matrices, which are high-order accurate, depending on the degree of the interpolat-

ing polynomial. Herein, we use quadratic hexahedral elements (27-noded) with a

Legendre-Gauss-Lobatto quadrature rule (Table 2.1).

Table 2.1: Legendre-Gauss-Lobatto quadrature rule.

Element Location of nodes Location of integration points Weights

Quadratic ±1.0 ±1.0 1/30.0 0.0 4/3

An m point Legendre-Gauss-Lobatto rule integrates polynomials of degree

up to and including 2m 3, exactly [77]. However, to compute mass-like matrices,

one needs to integrate terms with ΦΦT -like components, where Φ is the vector of

Lagrange interpolating polynomials (see A.1). Having m interpolation nodes results

in polynomials of degree m 1. The tensor products then involve terms of degree

2m 2; thus, the approach relies on under-integration in order to return a diagonal

mass-like matrix. Herein, we use the Legendre-Gauss-Lobatto rule to compute all

the submatrices presented in (2.31).

We remark that integration of mass-like matrices must be done consistently.

This means that the same quadrature rule must be used to compute MRD, Mi,Ni,

i = a, b, c, d in (2.31), therefore, rendering all these matrices diagonal. Choosing a

scheme that diagonalizes the mass-like matrix M in (2.31), whether done by conven-

tional mass lumping, or, via spectral elements, while not applying the same scheme

36

uniformly to all mass-like matrices, will result in instabilities, as it has also been

reported in [31, 79].

2.5 A symmetric formulation

In Section 2.2, we discussed a non-symmetric PML formulation that can be

integrated explicitly in time. In this section, we discuss an alternative formulation,

that results in a symmetric semi-discrete form, which would require an implicit time-

integration scheme due to a non-diagonal mass-like matrix. The key difference with

the Section 2.2 formulation is the handling of the combined constitutive and kine-

matic equations. To this end, we keep the equilibrium equation in (2.8a) intact, but

express (2.8b) in a different form.

Similar to what was done in [57], we start with the constitutive and kinematic

equations in the time domain

S = C[E], (2.37a)

E =1

2

[∇u+ (∇u)T

], (2.37b)

where E is the strain tensor, and C is the fourth-order constitutive tensor. For an

isotropic medium, C[E] = 2µE + λ(trE)I. Taking the Fourier transform of (2.37),

there results:

37

S = C[E], (2.38a)

E =1

2

[∇u+ (∇u)T

]. (2.38b)

Writing (2.38b) in the stretched coordinate system and using (2.4), we obtain:

E =1

2

(∇u)

1λx

0 0

0 1λy

0

0 0 1λz

+

1λx

0 0

0 1λy

0

0 0 1λz

(∇u)T

. (2.39)

Multiplying (2.39) by λxλyλz results in

λxλyλzE =1

2

[(∇u)Λ + Λ(∇u)T

]. (2.40)

Multiplying (2.40) by (iω)2 and using (2.13) and (2.2), we obtain

(iω)2aE+ iωbE+ cE+1

iωdE =

1

2(iω)2

[(∇u)Λe + Λe(∇u)T

]

+1

2iω

[(∇u)Λp + Λp(∇u)T

]+

1

2

[(∇u)Λw + Λw(∇u)T

]. (2.41)

Equation (2.41) constitutes the corresponding frequency-domain kinematic

equation in the stretched coordinate system. Taking the inverse Fourier transform

of (2.38a) and (2.41), there results

38

S = C[E], (2.42a)

aE+ bE + cE+ d

(∫ t

0

Edτ

)=

1

2

[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]. (2.42b)

Combining the resulting constitutive equation (2.42a) with the kinematic

equation (2.42b), and using the auxiliary variables introduced in (2.22), we obtain

D

[(a...S+ bS+ cS+ dS

)]=

1

2

[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

], (2.43)

where D is the compliance tensor (E = D[S]). Equation (2.43) constitutes the PML

combined constitutive and kinematic equations, which is equivalent to (2.23b).

Next, similar to what we did in Section 2.3.1, we take the inner product of

(2.43) with (tensor) test function T(x) ∈ L2(Ω); there results:

ΩPML

T :D(a...S+ bS+ cS+ dS

)dΩ

=1

2

ΩPML

T :[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]dΩ.

(2.44)

Upon discretization of (2.26a) and (2.44) via (2.28)-(2.29), we obtain a semi-discrete

form

39

Ms

...d+Csd+Ksd+Gsd = f , (2.45)

with the following definition for system matrices

Ms =

[MRD + Ma 0

0 −Na

], Cs =

[Mb Ae

ATe −Nb

], (2.46a)

Ks =

[KRD + Mc Ap

ATp −Nc

], Gs =

[Md Aw

ATw −Nd

], (2.46b)

d =[uh Sh

]T, f =

[fRD 0

]T, (2.46c)

where a bar denotes matrix extension to encompass all the displacement degrees-of-

freedom; Mi, i = a, b, c, d are PML matrices defined in (A.2b), and Ni, i = a, b, c, d,

Ai, i = e, p, w are defined in (A.7) and (A.8), respectively. Moreover, similar to

what we did in Section 2.3.2, (2.45) can be expressed similarly to (2.32) by taking

into account (2.32b), therefore, obviating the temporal differentiation of the forcing

term:

Msd+Csd+Ksd+Gsd = f . (2.47)

System matrices defined in (2.46a)-(2.46b) are now symmetric and indefinite

by contrast to (2.31a)-(2.31b). They can become positive definite if one multiplies

their lower blocks by a minus sign, at the expense of losing symmetry. We refer to

[33, 80] for details.

40

We remark that Na in M is a block penta-diagonal matrix; this entails an

implicit time-integration scheme for the semi-discrete form. The extended Newmark

method outlined in Appendix B.2 could then be used for time-stepping, which then

necessitates factorization of a symmetric matrix.

2.6 Generalization for multi-axial perfectly-matched-layers

The aforementioned derivations are based on using the classical stretching

function (2.2), where stretching is enforced only in the direction perpendicular to

the PML interface. It has been reported that, in two dimensions, and under cer-

tain parameterizations, this stretching function creates spurious growths when waves

travel along the interface, thus leading to numerical instability. In an attempt to sta-

bilize the PML (in 2D), Meza-Fajardo and Papageorgiou [41] proposed coordinate-

stretching in all directions within the PML buffer, leading to the, so-called, multi-

axial PML (M-PML).

Herein, we show that by making minimal modifications, our framework can

also accommodate the M-PML. We focus on the “right” PML buffer zone first, i.e.,

the volume contained in x0 ≤ x ≤ xt (see Fig. 2.1 with s replaced by x); extending

the ideas to the zones where two or three layers intersect is straightforward, and

can be accomplished by using superposition. We stretch the physical coordinates

according to

41

x = x0 +

∫ x

x0

[αx(x

′) +1

iωβx(x

′)

]dx′, (2.48a)

y = y0 +

∫ y

y0

[αy(x) +

1

iωβy(x)

]dy′, (2.48b)

z = z0 +

∫ z

z0

[αz(x) +

1

iωβz(x)

]dz′. (2.48c)

where αy, αz, βy, and βz are functions of x only, and are defined as

αy(x) = 1 + αo

[(x− xo)nx

LPML

]m, βy(x) = βx(x), (2.49a)

αz(x) = 1 + αo

[(x− xo)nx

LPML

]m, βz(x) = βx(x), (2.49b)

where is a proportionality constant, and nx is the outward unit normal at the

interface, similar to ns in Fig. 2.1; αx and βx are defined in (2.5). We remark that

αy, αz would have been reduced to one, and βy, βz would have been identically zero,

in the right buffer, had we used the classical stretching. Applying the fundamental

theorem of calculus to (2.48), results in

λx :=∂x

∂x= αx(x) +

1

iωβx(x), (2.50a)

λy :=∂y

∂y= αy(x) +

1

iωβy(x), (2.50b)

λz :=∂z

∂z= αz(x) +

1

iωβz(x). (2.50c)

These are the stretching functions the authors used in [45]. However, the definition

of the stretched gradient operator in equation (3) in [45] requires additional terms,

42

which the authors had not included. For example, the derivative with respect to x

should read

∂( )

∂x=

∂( )

∂x

∂x

∂x+

∂( )

∂y

∂y

∂x+

∂( )

∂z

∂z

∂x(2.51a)

=1

λx

∂( )

∂x− λyx

λxλy

∂( )

∂y− λzx

λxλz

∂( )

∂z, (2.51b)

instead of the expression given in [45], which reads

∂( )

∂x=

1

λx

∂( )

∂x. (2.52a)

In (2.51) above, the cross-derivative terms are defined as

λyx :=∂y

∂x=

(∂

∂xαy(x) +

1

∂xβy(x)

)(y − y0), (2.53a)

λzx :=∂z

∂x=

(∂

∂xαz(x) +

1

∂xβz(x)

)(z − z0). (2.53b)

In other words, it seems that in [41, 45], the authors have not accounted prop-

erly for the Jacobian. Thus, there are at least two possible forms of the M-PML: the

uncorrected form in [41, 45], and the corrected form, which accounts for the cross-

derivatives. Interestingly, numerical experiments we performed in two dimensions

with the corrected form yielded small, but non-negligible reflections from the inter-

face. By contrast, the uncorrected form yielded better results, despite its unsound

mathematical foundation. This has led us to adopt the approach taken in [41, 45];

accordingly, the equation pertaining to the conservation of linear momentum in the

stretched coordinate system becomes

43

div(STΛ

)− ST divΛ = (iω)2ρ λxλyλz u, (2.54)

which results in the following strong form

div(STΛe + STΛp + STΛw

)−

(ST divΛe + ST divΛp + ST divΛw

)

= ρ (a...u+ bu+ cu+ du) . (2.55)

The structure of the formulation pertaining to the combined constitutive and

kinematic equation in the stretched coordinate system we previously discussed re-

mains unaltered. Hence, for accommodating the M-PML, one only needs to replace

(2.24b) in the strong form of the equations with (2.55), which, in turn, changes the

definition of submatrices Aeu,Apu, and Awu in (2.31). The new definition of these

submatrices for the M-PML case are given in Appendix A.3.

2.7 Numerical Experiments

We present three numerical experiments to test the accuracy and efficacy

of our hybrid formulation. The first example involves a homogeneous half-space;

the second one is a horizontally layered medium with an ellipsoidal inclusion. The

last example compares various formulations discussed in Sections 2.3.2 and 2.5. We

compare our results against an enlarged domain6 solution with fixed boundaries,

obtained via a standard displacement-only formulation, which may be viewed as

6denoted by ΩED, such that ΩRD ⊂ ΩED.

44

a reference solution. Due to the fixed boundaries of the enlarged domain model,

reflection occurs at these boundaries; hence, we limit the comparison time up to the

arrival of the reflected waves to the regular domain.

In the first two examples, we apply a surface traction on the medium, with a

Ricker pulse time signature, defined as

Tp(t) =(0.25u2 − 0.5)e−0.25u2 − 13e−13.5

0.5 + 13e−13.5with 0 ≤ t ≤ 6

√6

ωr, (2.56)

such that

u = ωrt− 3√6, (2.57)

where ωr (= 2πfr) denotes the characteristic central circular frequency of the pulse.

Here, we take fr = 15 Hz, and the load has an amplitude of 1 kPa. The pulse

time-history, and its corresponding Fourier spectrum are shown in Fig. 2.4.

0 0.1 0.2 0.3 0.4 0.5−1

−0.5

0

0.5

t (sec)

f(k

Pa)

0 15 30 45 600

1

2

3x 10

−3

frequency (Hz)

|f(ω

)|

Figure 2.4: Ricker pulse time history and its Fourier spectrum.

45

In order to quantify the performance of our PML formulation, we consider two met-

rics: (a) time-history comparisons at select nodes via evaluation of a time-dependent

Euclidean norm of the error, relative to the reference solution; and (b) decay of the

total energy within the regular domain.

We define the time-dependent Euclidean norm of the relative error at a point

x ∈ ΩRD as

e(x, t) =‖u(x, t)− uED(x, t)‖2

maxt‖uED(x, t)‖2

, (2.58)

where uED(x, t) represents the enlarged domain solution, and the Euclidean norm of

a vector u(x, t) = [ux(x, t), uy(x, t), uz(x, t)]T is defined as

‖u(x, t)‖2 =√(

ux(x, t))2

+(uy(x, t)

)2+(uz(x, t)

)2. (2.59)

The energy introduced into the system through the loading is carried via

waves, which then enter the PML buffer and attenuate. Therefore, an effective PML

ought to result in the rapid decay of energy. The total energy of the system can be

computed at any time via

Et(t) =1

2

ΩRD

ρ(x)[uT (x, t)u(x, t)

]dΩ +

1

2

ΩRD

[σT (x, t)ǫ(x, t)

]dΩ, (2.60)

where u denotes the velocity vector, and σ and ǫ are stress and strain vectors, respec-

tively. We compute and compare energy only within the regular domain. Moreover,

46

Et(t) can also be used as a stability indicator since one expects that the total energy

decays monotonically for a stable formulation.

2.7.1 Homogeneous media

We consider a homogeneous half-space with shear wave velocity cs = 500 m/s,

Poisson’s ratio ν = 0.25, and mass density ρ = 2000 kg/m3, which, after truncation,

is reduced to a cubic computational domain of length and width 100 m × 100 m,

and 50 m depth. A 12.5 m-thick PML is placed at the truncation boundaries, as

shown in Fig. 2.5. Two excitations are considered: a vertical stress load (vertical

excitation), and a horizontal traction along the x axis (horizontal excitation). The

excitations have the Ricker pulse temporal variation (Fig. 2.4), and are applied on

the surface of the medium over a region (−1.25 m ≤ x, y ≤ 1.25 m). We carry

out the simulation for each excitation separately. The interior and PML domains

are discretized by quadratic hexahedral spectral elements (i.e., 27-noded bricks, and

quadratic-quadratic pairs of approximation for displacement and stress components

in the PML) of size 1.25 m. For the PML parameters, we choose αo = 5, βo = 866 s 1,

and a quadratic profile for the attenuation functions, i.e., m = 2. Using the 4th-order

explicit Runge-Kutta method, discussed in B.1, with a time step of ∆t = 0.0006 s, we

compute the response for 2 s using the hybrid formulation corresponding to (2.36).

We also compute a reference solution, via a standard displacement-only for-

mulation, for an enlarged domain of size 440 m × 440 m × 220 m, with fixed

boundaries, using the same element type and size discussed above. For this example,

P-wave velocity is cp = 866 m/s. Therefore, it takes 0.45 s for the P-wave generated

47

100m100m

x y

z

load

50m

Figure 2.5: PML-truncated semi-infinite homogeneous media.

by the stress load, which is applied at the center of the surface, to hit the fixed

boundaries and return to the regular domain. We use the 4th-order explicit Runge-

Kutta method, with ∆t = 0.0006 s for time-stepping and compute the response for

0.45 s. Table 2.2 summarizes the discretization details of the two considered models7.

Table 2.2: Discretization details of the hybrid-PML and enlarged domain models.

elements nodes unknowns

hybrid-PML 500,000 4,080,501 24,228,426enlarged-domain 21,807,104 175,449,825 521,884,704

Figure 2.6 displays snapshots of the total displacement at two different times

for the vertical excitation. The left figure shows waves at an evolving stage, while the

figure on the right demonstrates absorption of waves in the PML region. Figure 2.7

shows the corresponding wave motion for the horizontal excitation. No discernible

7We developed a code in Fortran, using PETSc [81] to facilitate parallel implementation.

48

reflections can be observed from the PML interface, nor any residuals from the fixed-

end boundaries, indicating satisfactory performance of the PML.

Figure 2.6: Snapshots of total displacement taken at t = 0.111 s, 0.219 s (verticalexcitation).

We compare time histories of the hybrid PML formulation against the refer-

ence solution at select points. The location of these points are summarized in Table

2.3; the maximum relative error at each of these sampling points, computed using

(2.58), is presented in the fifth and the sixth column, for the vertical and horizon-

tal excitations, respectively. The relative error is very small and demonstrates the

efficacy and success of the approach.

Figures 2.8 and 2.9 display comparison of the two responses, due to the vertical

and horizontal excitation, at various sampling points. The agreement is excellent;

the PML has effectively absorbed waves with practically no reflections. The response

is causal, effectively dies out at around t = 0.35 s at all the considered points, and

is free from spurious reflections.

49

0 6e-5

2e-5 4e-5

Figure 2.7: Snapshots of total displacement taken at t = 0.147 s, 0.219 s (horizontalexcitation).

Table 2.3: Relative error at sampling points between hybrid-PML and enlarged do-main solutions.sample x y z error (homogeneous) error (homogeneous) error (heterogeneous)point vertical excitation horizontal excitation vertical excitation

sp1 0 0 0 1.17× 10−12 8.77× 10−13 4.61× 10−10

sp2 +50 0 0 2.52× 10−8 2.34× 10−8 6.07× 10−7

sp3 +50 0 -25 2.89× 10−9 4.12× 10−8 2.87× 10−6

sp4 +50 0 -50 1.46× 10−7 1.42× 10−7 7.03× 10−6

sp5 0 0 -50 9.86× 10−9 2.51× 10−9 1.41× 10−5

sp6 +50 +50 0 3.26× 10−7 2.16× 10−7 1.86× 10−6

sp7 +50 +50 -25 5.50× 10−8 1.18× 10−7 6.72× 10−6

sp8 +50 +50 -50 5.08× 10−7 5.25× 10−7 6.44× 10−6

50

0 0.1 0.2 0.3 0.4 0.5−5

−2.5

0

2.5

5x 10

−5

t (sec)

ux

(m)

sp2

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−1.5

−0.75

0

0.75

1.5x 10

−5

t (sec)

uz

(m)

sp4

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−2.6

−1.3

0

1.3

2.6x 10

−5

t (sec)

ux

(m)

sp6

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−1.5

−0.75

0

0.75

1.5x 10

−5

t (sec)

uz

(m)

sp8

Enlarged domainPML-truncated domain

Figure 2.8: Comparison of displacement time histories between the enlarged andPML-truncated domain solutions at the sp2, sp4, sp6, and sp8 sampling points (ho-mogeneous case, vertical excitation).

51

0 0.1 0.2 0.3 0.4 0.5−3

−1.5

0

1.5

3x 10

−5

t (sec)

ux

(m)

sp2

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−3

0

3

6x 10

−5

t (sec)

uz

(m)

sp2

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−1

0

1

2x 10

−5

t (sec)

uy

(m)

sp8

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

4x 10

−6

t (sec)

uz

(m)

sp8

Enlarged domainPML-truncated domain

Figure 2.9: Comparison of displacement time histories between the enlarged andPML-truncated domain solutions at the sp2 and sp8 sampling points (homogeneouscase, horizontal excitation).

52

Figure 2.10 shows the normalized error time history (2.58) due to the vertical

excitation for two distinct locations: sp3 and sp8. Figure 2.11 shows the correspond-

ing error time history for the horizontal excitation. Among all the considered loca-

tions, sp8 has the highest error, which is only 5.08× 10−7 for the vertical excitation,

and 5.25× 10−7 for the horizontal excitation.

0 0.1 0.2 0.3 0.4 0.50

1

2

3x 10

−9

t (sec)

norm

alize

der

ror

sp3

0 0.1 0.2 0.3 0.4 0.50

2

4

6x 10

−7

t (sec)

norm

alize

der

ror

sp8

Figure 2.10: Relative error time history e(x, t) at various sampling points (homoge-neous case, vertical excitation).

The total energy decay within the regular domain, due to the vertical ex-

citation, is plotted in Figure 2.12, both in standard and semi-logarithmic scale for

various values of βo. Due to the limited size of the enlarged domain model, we com-

pare the enlarged domain solution with the set of the PML solutions for various βo

only up to 0.45 s, since for t > 0.45 s, the reflections from the enlarged domain’s

fixed boundaries would have traveled back to the regular domain. The agreement is

excellent and no difference can be observed. For βo = 866 s 1, the total reduction in

53

0 0.1 0.2 0.3 0.4 0.50

2

4

6x 10

−8

t (sec)

norm

alize

der

ror

sp3

0 0.1 0.2 0.3 0.4 0.50

2

4

6x 10

−7

t (sec)norm

alize

der

ror

sp8

Figure 2.11: Relative error time history e(x, t) at various sampling points (homoge-neous case, horizontal excitation).

energy, relative to its peak value, is 14 orders of magnitude. The decay is sharp and

smooth, without any discernible reflections, indicating the effectiveness and health

of the PML. Figure 2.13 displays the corresponding decay of energy curves due to

the horizontal excitation. As can be seen from the standard scale plot, most of the

energy travels out of the interior domain quickly, and gets absorbed in the PML

effectively.

To illustrate the stability of the formulation, we run the simulation for 50,000

time steps. The total energy decay is displayed in Fig. 2.14 and shows no numerical

instability during the simulation time.

We remark that we also used M-PML terminations to conduct this numerical

experiment. While the results are satisfactory in general, they are not as accurate

as when using PML terminations. In fact, for all the sampling points of Table 2.3,

54

0 0.5 1 1.5 20

1000

2000

3000

4000

5000

6000

7000

8000

t (sec)

Et(J

)

Enld. dom.βo = 700βo = 800βo = 866βo = 900βo = 1000

(a) standard scale

0 0.5 1 1.5 210

−15

10−10

10−5

100

105

t (sec)E

t(J

)

Enld. dom.βo = 700βo = 800βo = 866βo = 900βo = 1000

(b) logarithmic scale

Figure 2.12: Total decay of energy within the regular domain for various values ofβo (homogeneous case, vertical excitation).

0 0.5 1 1.5 20

1000

2000

3000

4000

5000

6000

7000

8000

9000

t (sec)

Et(J

)

Enld. dom.βo = 700βo = 800βo = 866βo = 900βo = 1000

(a) standard scale

0 0.5 1 1.5 210

−15

10−10

10−5

100

105

t (sec)

Et(J

)

Enld. dom.βo = 700βo = 800βo = 866βo = 900βo = 1000

(b) logarithmic scale

Figure 2.13: Total decay of energy within the regular domain for various values ofβo (homogeneous case, horizontal excitation).

55

0 10 20 3010

−15

10−10

10−5

100

105

t (sec)

Et(J

)

(a) vertical excitation

0 10 20 3010

−15

10−10

10−5

100

105

t (sec)E

t(J

)

(b) horizontal excitation

Figure 2.14: Total decay of energy within the regular domain for βo = 866 s 1

(homogeneous case).

the relative error remained less than 1%, except at the corner point sp8 where the

relative error is about 5%. We do not report the M-PML-based results in detail

pending comprehensive investigations.

2.7.2 Heterogeneous media

In the second example, we study the performance of our hybrid-PML formu-

lation for a heterogeneous medium. We consider a 100 m × 100 m × 50 m layered

medium with an ellipsoidal inclusion, where a 12.5 m-thick PML is placed at its

truncation boundaries, as shown in Fig. 2.15. The properties of the medium are

56

cs(z) =

400 m/s, for − 20 m ≤ z ≤ 0 m,

500 m/s, for − 50 m ≤ z < −20 m,

600 m/s, for ellipsoidal inclusion,

(2.61)

with mass density ρ = 2000 kg/m3, and Poisson’s ratio ν = 0.25; the ellipsoidal

inclusion occupies the region (x−2515

)2+(y−255

)2+( z+205

)2 ≤ 1. The material properties

at the interfaces ΓI are extended into the PML buffer. A vertical stress load, with

the Ricker pulse temporal signature, is applied on the surface of the medium over

a region (−1.25 m ≤ x, y ≤ 1.25 m). The problem is discretized with quadratic

hexahedral elements of size 1.25 m. The PML parameters are taken as αo = 5,

βo = 500 s 1, and m = 2. To resolve the motion we use the first-order system (2.36),

and a 4th-order explicit Runge-Kutta method, with a time step of ∆t = 0.00048 s.

100m10

0m

x y

z

load

20m

30m

30m

Figure 2.15: PML-truncated semi-infinite heterogeneous media.

Next, a reference solution is computed by using an enlarged domain, with

size and spatial discretization properties similar to the previous example, and a time

57

step of ∆t = 0.00048 s. Table 2.2 summarizes the spatial discretization details of the

problem.

To assess the performance of our hybrid PML formulation, we compare dis-

placement time histories at select locations against the enlarged domain solution.

The sampling points, with their corresponding relative error computed via (2.58) are

summarized in the last column of Table 2.3. The relative error values are higher than

the previous example, which was a homogeneous medium, but still, they are very

low. In fact, the highest relative error, which corresponds to sp8, is only 1.41×10−5,

which is very small in practical applications.

Snapshots of the total displacement at two different times are displayed in

Fig. 2.16. The figure on the left shows waves at an evolving stage, while the right

figure indicates absorption of waves in the PML buffer zone. Notice that no dis-

cernible reflections can be seen from the PML interfaces, nor any residuals from the

fixed-end boundaries, a visual indication of satisfactory performance of the PML

for domains involving heterogeneous material properties. Figure 2.17 depicts the

complex wave pattern around the ellipsoidal inclusion on a cross-section through the

domain situated at 20 m from the surface going through the ellipsoid’s midplane.

Various components of displacement time-histories for the enlarged domain

and the hybrid PML simulations are displayed in Fig. 2.18 for select sampling points.

The agreement is excellent. The response effectively dies out at around 0.45 s.

The relative error time histories are shown in Fig. 2.19 and indicate satisfactory

performance of the PML.

58

Figure 2.16: Snapshots of total displacement taken at t = 0.111 s, 0.225 s.

3.0e-06 3.0e-05

1.0e-05

1.0e-06 1.0e-05

Figure 2.17: Snapshots of total displacement taken at t = 0.233 s, and 0.290 s, onthe z = −20 m domain cross-section.

59

0 0.1 0.2 0.3 0.4 0.5−6

−3

0

3

6

9

12x 10

−6

t (sec)

ux

(m)

sp3

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

1.5x 10

−5

t (sec)

uz

(m)

sp5

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

4

6x 10

−5

t (sec)

uy

(m)

sp6

Enlarged domainPML-truncated domain

0 0.1 0.2 0.3 0.4 0.5−1.6

−0.8

0

0.8

1.6

2.4x 10

−5

t (sec)

uz

(m)

sp7

Enlarged domainPML-truncated domain

Figure 2.18: Comparison of displacement time histories between the enlarged andPML-truncated domain solutions at the sp3, sp5, sp6, and sp7 sampling points (het-erogeneous case).

60

0 0.1 0.2 0.3 0.4 0.50

0.5

1

1.5x 10

−5

t (sec)

norm

alize

der

ror

sp5

0 0.1 0.2 0.3 0.4 0.50

2

4

6

8x 10

−6

t (sec)norm

alize

der

ror

sp8

Figure 2.19: Relative error time history at various sampling points (heterogeneouscase).

We compare the total energy decay within the regular domain between the

PML and enlarged domain solutions. Due to the size of the enlarged domain model,

the energy can only be computed up to 0.45 s. For the PML-truncated model,

however, we allow the simulation to run for 2 s. The energy curves are presented

in Fig. 2.20. The agreement between the enlarged domain solution and the PML

curves is excellent. The reduction of the total energy is 14 orders of magnitude for

βo = 500 s 1; energy decay is sharp and smooth, with no sign of reflections, signifying

satisfactory performance of the PML.

Finally, to illustrate the stability of the proposed scheme, we run the simula-

tion for 125,000 time steps. The decay of the total energy is shown in Fig. 2.21 and

shows no numerical instability during the simulation time.

61

0 0.5 1 1.5 20

2000

4000

6000

8000

10000

12000

14000

t (sec)

Et(J

)

Enld. dom.βo = 400βo = 500βo = 600

(a) normal scale

0 0.5 1 1.5 210

−15

10−10

10−5

100

105

t (sec)E

t(J

)

Enld. dom.βo = 400βo = 500βo = 600

(b) logarithmic scale

Figure 2.20: Total decay of energy within the regular domain for various values ofβo (heterogeneous case).

0 10 20 30 40 50 6010

−15

10−10

10−5

100

105

t (sec)

Et(J

)

Figure 2.21: Total decay of energy within the regular domain for βo = 500 s 1

(heterogeneous case).

62

2.7.3 Comparison of various formulations

In the previous numerical experiments, we utilized explicit time-stepping us-

ing spectral elements, which is well suited for solving large-scale problems on parallel

computers. In this section, we compare alternative formulations and various time

marching schemes, discussed earlier in Sections 2.3.2 and 2.5, using the same numeri-

cal experiment considered in [31]. Specifically, we use the standard Newmark method

for the second-order in time forms (2.34, 2.35) using standard quadratic 20-noded

elements. Next, we apply the extended Newmark method to the symmetric form

(2.47) also discretized by standard quadratic 20-noded elements. We also compute

the response of the system with explicit RK-4 scheme for both PML-truncated, and

an enlarged domain using quadratic spectral elements. The size of the considered

enlarged domain model allows simulation for up to 15 s, before reflections travel

back to the interior domain. Results for PML-truncated domain models, however,

are computed for 30 s. The numerical experiments are summarized in Table 2.4.

Table 2.4: Test cases for comparing various formulations and their correspondingrelative error.Case Equation Time-stepping Element type error (center) error (corner)

UnSym-1 (2.34)-2nd order standard Newmark Lagrangian 20-noded 1.73× 10−5 2.31× 10−4

UnSym-2 (2.35)-2nd order standard Newmark Lagrangian 20-noded 1.73× 10−5 2.31× 10−4

Symmetric (2.47)-3rd order extended Newmark Lagrangian 20-noded 1.73× 10−5 2.31× 10−4

Explicit (2.36)-1st order RK-4 Spectral 27-noded 5.03× 10−6 1.80× 10−5

Enld. dom. 1st order RK-4 Spectral 27-noded(no PML)

The problem consists of a half-space with shear wave velocity cs = 1 m/s,

Poisson’s ratio ν = 0.25, and mass density ρ = 1 kg/m3, which, after truncation, is

63

reduced to a 1.2 m× 1.2 m× 0.2 m regular domain, and 0.8 m-thick PML is placed

on the sides and at the bottom of the truncation boundaries, as shown in Fig. 2.22.

A uniform pressure, of the form considered in [2] (with characteristic parameters

td = 10 s, ωf = 3 rad/s), as shown in Fig. 2.23, is applied on the surface over

a region (−1 m ≤ x, y ≤ 1 m). Quadratic elements of size 0.2 m are used for

discretizing both the interior domain and the PML buffer (i.e., quadratic-quadratic

pairs of approximation for displacement and stress components in the PML). We

consider αo = 10, βo = 20 s 1, and m = 2 for the PML parameters, and ∆t = 0.05 s

for temporal discretization.

load

Figure 2.22: Quarter model of a PML-truncated semi-infinite homogeneous media.

Considering the enlarged domain solution as benchmark, we compute the

maximum relative error given by (2.58) at the center and corner of the loading

surface. These values are given in Table 2.4, and are very small, considering that

PML was discretized only with four elements. The largest relative error is only

2.31 × 10−4. The relative error for the cases using Newmark schemes are slightly

greater than those using RK-4, as one would expect. The vertical component of

64

0 5 10 15 20−1.5

−1

−0.5

0

0.5

1

1.5

t (sec)

f(P

a)

0 2 4 6 8 100

0.02

0.04

0.06

frequency (rad/s)

|f(ω

)|

Figure 2.23: Surface load time history considered in Section 2.7.3 and its Fourierspectrum.

the displacement time history for the center and corner nodes is depicted in Fig.

2.24; the agreement is remarkable. Overall, all cases considered in Table 2.4 provide

satisfactory results.

0 10 20 30−0.6

−0.3

0

0.3

0.6

t (sec)

uz

(m)

center

UnSym-1UnSym-2SymmetricExplicitEnld. dom.

0 10 20 30−0.2

−0.1

0

0.1

0.2

t (sec)

uz

(m)

corner

UnSym-1UnSym-2SymmetricExplicitEnld. dom.

Figure 2.24: Comparison of displacement time histories for various cases consideredin Table 2.4.

65

2.8 Summary

We presented various time-domain formulations for elastic wave propagation

in arbitrarily heterogeneous PML-truncated media. The main contribution of this

development over the earlier work [33] is the extension to three-dimensions of a

hybrid formulation, endowed with explicit time integration and spectral elements.

Our formulation is hybrid in the sense that it uses a displacement-stress formulation

for the PML buffer, coupled with a standard displacement-only formulation for the

interior domain, which results in optimal computational cost and allows for the

ready incorporation of the PML in existing codes. Moreover, the mixed-field finite

element scheme for the PML buffer does not require specialized elements for LBB-

type stability.

The resulting semi-discrete form of the PML-truncated model is third-order-

in-time. Several alternatives for time marching were discussed which may suit various

applications. In particular, we discussed: (a) an explicit time stepping scheme utiliz-

ing the Runge-Kutta method; (b) time integration via the standard Newmark scheme

by recasting the semi-discrete form into a second-order system; and (c) applying an

extended Newmark scheme to a fully-symmetric, third-order-in-time, semi-discrete

form. Numerical experiments demonstrate stability, efficacy, and satisfactory perfor-

mance of the proposed schemes.

66

Chapter 3

The elastic inverse medium problem in

three-dimensional PML-truncated domains

In this chapter, we address the inverse medium problem of finding the dis-

tributed elastic properties of an arbitrarily heterogeneous soil medium. To allow for

this chapter to be self-contained, we discuss first the forward scheme we chose to use

among those presented in the preceding chapter. Next, we discuss the mathematical

and numerical aspects of the underlying inverse medium problem, where we derive

the adjoint and control problems, and discuss strategies that invite robustness. We

report on numerical experiments, using synthetic data, targeting the reconstruction

of both smooth and sharp profiles. Lastly, we conclude with summary remarks.

3.1 The forward problem

Consider the forward problem first cast in (2.24)-(2.25). After integrating the

equations once in time, there result:

67

divµ[∇u+ (∇u)T

]+ λ(divu)I

+ b = ρu in ΩRD × J,

(3.1a)

div(STΛe + STΛp + STΛw

)= ρ (au+ bu+ cu+ du) in ΩPML × J,

(3.1b)

aS+ bS+ cS+ dS =

µ[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]+

λ [div(Λeu) + div(Λpu) + div(Λwu)] I in ΩPML × J.(3.1c)

The system is initially at rest, and subject to the following boundary and interface

conditions:

µ[∇u+ (∇u)T

]+ λ(divu)I

n = gn on ΓRD

N × J, (3.2a)

(STΛe + STΛp + STΛw)n = 0 on ΓPMLN × J,

(3.2b)

u = 0 on ΓPMLD × J,

(3.2c)

uRD = uPML on ΓI × J, (3.2d)

µ[∇u+ (∇u)T

]+ λ(divu)I

n = (STΛe + STΛp + STΛw)n on ΓI × J, (3.2e)

where again, a bar ( ) indicates history of the subtended variable1.

Next, we seek a weak solution, corresponding to the strong form of (3.1)

and (3.2), in the Galerkin sense, similarly to the steps followed in Section 2.3.1.

1For instance, u(x, t) =∫ t

0u(x, τ) dτ .

68

Specifically, we take the inner products of (3.1a) and (3.1b) with (vector) test function

w(x), and integrate by parts over their corresponding domains. Incorporating (3.2d-

3.2e) eliminates the interface boundary terms and results in (3.3a). Next, we take

the inner product of (3.1c) with (tensor) test function T(x); there results (3.3b).

Accordingly, find u ∈ H1(Ω)× J, and S ∈ L2(Ω)× J, such that:

ΩRD

∇w :µ[∇u+ (∇u)T

]+ λ(divu)I

dΩ +

ΩPML

∇w :(STΛe + STΛp + STΛw

)dΩ

+

ΩRD

w · ρu dΩ +

ΩPML

w · ρ (au+ bu+ cu+ du) dΩ =

ΓRD

N

w · gn dΓ +

ΩRD

w · b dΩ,

(3.3a)

ΩPML

T :(aS+ bS+ cS+ dS

)dΩ

=

ΩPML

T :µ[(∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw + Λw(∇u)T

]

+ T :λ [div(Λeu) + div(Λpu) + div(Λwu)] I dΩ, (3.3b)

for every w ∈ H1(Ω) and T ∈ L2(Ω), where gn ∈ L2(Ω)× J, and b ∈ L2(Ω)× J.

Then the following semi-discrete form results:

Mdst +Cdst +Kdst +Gdst = f st, (3.4a)

dst =

∫ t

0

dst(τ)|PML dτ, (3.4b)

69

where spatial and temporal dependencies are suppressed for brevity; M, C, K, G,

are system matrices2, dst = (uTh ,S

Th )

T is the vector of nodal unknowns comprising

displacements in ΩRD ∪ ΩPML, and stress components only in ΩPML, and f st is the

vector of applied forces. Equations (3.4) are identical to (2.32), which were derived

following a different sequence of temporal operations. The superscript st is used here

to designate the solution of the forward problem as the “state” solution, and contrast

it with the solution of the adjoint problem, which is discussed in Section 3.2.1.2.

The matrix M has a block-diagonal structure, and can be diagonalized if one

uses spectral elements, which then enables the explicit time integration of (3.4). In

this regard, we express (3.4) as a first-order system

d

dt

x0

x1

Mx2

=

0 I 00 0 I−G −K −C

x0

x1

x2

+

00f st

, (3.5)

where x0 = dst, x1 = dst, and x2 = dst. We then use an explicit fourth-order

Runge-Kutta (RK-4) method for integrating (3.5) in time.

3.2 The inverse medium problem

Our goal is to find the distribution of the Lame parameters λ(x) and µ(x)

within the elastic soil medium. We consider the sources, and the response recorded

at receivers on the ground surface, as known. The inverse medium problem can thus

be formulated as the minimization of the difference (or misfit) between the measured

2The definition of M, C, K, and G are given in (2.31).

70

response at receiver locations, and a computed response corresponding to a trial

distribution of the material parameters. The misfit minimization should honor the

physics of the problem, as described by the forward problem of the preceding section.

Mathematically, the inverse medium problem can be cast as a PDE-constrained

optimization problem:

minλ,µ

J(λ, µ) :=1

2

Nr∑

j=1

∫ T

0

Γm

(u− um) · (u− um) δ(x− xj) dΓ dt+ R(λ, µ), (3.6)

where the minimization is constrained by the forward initial and boundary value

problem (PDE-constrained). In (3.6), u is the solution of the forward problem gov-

erned by the initial- and boundary-value problem (3.1), (3.2), J is the objective

functional3, Nr denotes the total number of receivers, T is the total observation

time, Γm is the part of the ground surface where the receiver response, um, has

been recorded, δ(x − xj) is the Dirac delta function, which enables measurements

at receiver locations xj , and R(λ, µ) is the regularization term, which is discussed

below.

Inverse problems suffer from solution multiplicity, which, in general, is due to

the presence of insufficient data. This makes the problem ill-posed in the Hadamard

sense. Regularization of the solution by using the Tikhonov (TN) [84], or, the Total

Variation (TV) [85] scheme are among common strategies to alleviate ill-posedness.

3We use J to indicate the corresponding discrete objective functional. See [82, 83] for otherpossibilities.

71

The Tikhonov regularization, denoted by RTN(λ, µ), penalizes large material gradi-

ents and, thus, precludes spatially rapid material variations from becoming solutions

to the inverse medium problem. It is defined as:

RTN(λ, µ) =

2

ΩRD

∇λ · ∇λ dΩ +Rµ

2

ΩRD

∇µ · ∇µ dΩ, (3.7)

where Rλ and Rµ are the so-called λ- and µ-regularization factor, respectively, and

control the amount of penalty imposed via (3.7) on the gradients of λ and µ. By

construction, TN regularization results in material reconstructions with smooth vari-

ations. Consequently, sharp interfaces may not be captured well when using the TN

scheme. The TV regularization, however, works better for imaging profiles involving

sharp interfaces, as it typically preserves edges. It is defined as:

RTV (λ, µ) =Rλ

2

ΩRD

(∇λ · ∇λ+ ǫ)1

2 dΩ +Rµ

2

ΩRD

(∇µ · ∇µ+ ǫ)1

2 dΩ, (3.8)

where the parameter ǫ makes RTV differentiable when either ∇λ · ∇λ = 0, or,

∇ µ · ∇ µ = 0.

For computing the first-order optimality conditions for (3.6), we use the (for-

mal) Lagrangian approach [86] to impose the PDE constraint in its weak form. These

are necessary conditions that must be satisfied at a stationary point of (3.6). Specif-

ically, we introduce Lagrange multiplier vector function w ∈ H1(Ω), and Lagrange

multiplier tensor function T ∈ L2(Ω) to enforce the initial- and boundary-value

72

problem (3.1), (3.2), which admits the weak form (3.3). The Lagrangian functional

becomes

L(u,S,w,T, λ, µ) =1

2

Nr∑

j=1

∫ T

0

Γm

(u− um) · (u− um) δ(x− xj) dΓ dt+ R(λ, µ)

−∫ T

0

ΩRD

∇w :µ[∇u+ (∇u)T

]+ λ(divu)I

dΩ dt

−∫ T

0

ΩPML

∇w :(STΛe + STΛp + STΛw

)dΩ dt−

∫ T

0

ΩRD

w · ρu dΩ dt

−∫ T

0

ΩPML

w · ρ (au+ bu+ cu+ du) dΩ dt+

∫ T

0

ΓRD

N

w · gn dΓ dt

+

∫ T

0

ΩRD

w · b dΩ dt−∫ T

0

ΩPML

T :(aS+ bS+ cS+ dS

)dΩ dt

+

∫ T

0

ΩPML

T :µ((∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T + (∇u)Λw

+ Λw(∇u)T)+T :λ [div(Λeu) + div(Λpu) + div(Λwu)] I dΩ dt, (3.9)

where now u, S, λ, and µ are treated as independent variables.

3.2.1 Optimality system

We use the Lagrangian functional (3.9) as a tool to compute the optimality

system for (3.6). To this end, the Gateaux derivative4 (or first variation) of the

Lagrangian functional with respect to all variables must vanish. This process is

4See Appendix C for the definition and notation.

73

discussed next.

3.2.1.1 The state problem

Taking the derivatives of the Lagrangian functional L with respect to w and

T in directions w ∈ H1(Ω) and T ∈ L2(Ω), and setting it to zero, results in the state

problem, which is identical to (3.3). That is

L′(u,S,w,T, λ, µ)(w, T) = 0. (3.10)

3.2.1.2 The adjoint problem

We now take the derivative of L with respect to u and S in directions

u ∈ H1(Ω) and S ∈ L2(Ω). This yields

74

L′(u,S,w,T, λ, µ)(u, S) =Nr∑

j=1

∫ T

0

Γm

u · (u− um) δ(x− xj) dΓ dt

−∫ T

0

ΩRD

∇w :µ[∇u+ (∇u)T

]+ λ(div u)I

dΩ dt

−∫ T

0

ΩPML

∇w :(˙STΛe + STΛp +

¯STΛw

)dΩ dt−

∫ T

0

ΩRD

w · ρ¨u dΩ dt

−∫ T

0

ΩPML

w · ρ(a¨u+ b ˙u+ cu+ d ¯u

)dΩ dt

−∫ T

0

ΩPML

T :(a¨S+ b ˙S+ cS+ d ¯S

)dΩ dt

+

∫ T

0

ΩPML

T :µ((∇ ˙u)Λe + Λe(∇ ˙u)T + (∇u)Λp + Λp(∇u)T + (∇¯u)Λw

+ Λw(∇¯u)T)+T :λ

[div(Λe

˙u) + div(Λpu) + div(Λw¯u)]I dΩ dt. (3.11)

Setting the above derivative to zero, and performing integration by parts in time,

results in the statement of the weak form of the adjoint problem. That is, find

w ∈ H1(Ω)× J, and T ∈ L2(Ω)× J, such that:

75

ΩRD

∇u :µ[∇w + (∇w)T

]+ λ(divw)I

dΩ+

+

ΩRD

u · ρw dΩ +

ΩPML

u · ρ (aw − bw + cw− dw) dΩ

−∫

ΩPML

∇u :µ[−TΛe − TTΛe +TΛp +TTΛp − TΛw − TTΛw

]

+ λ[−T : div(Λeu) +T : div(Λpu)− T : div(Λwu)

]I dΩ

=

Nr∑

j=1

Γm

u · (u− um) δ(x− xj) dΓ, (3.12a)

ΩPML

∇w : STΛe −∇w : STΛp +∇w : STΛw dΩ =

ΩPML

S :(aT− bT + cS− dT

)dΩ

(3.12b)

for every u ∈ H1(Ω) and S ∈ L2(Ω), where w(x, T ) = 0, and T(x, T ) = 0.

We remark that the adjoint problem (3.12) is a final-value problem and, thus,

is solved backwards in time5; it is driven by the misfit between a computed response,

and the measured response at receiver locations. Moreover, the operators implicated

in the adjoint problem are very similar to those of the state problem: they involve

transposition of the system matrices, and sign reversal for terms involving history,

and first-order time derivatives. In this regard, we obtain the following semi-discrete

form for the adjoint problem:

5See [87] for other possibilities, and [88–90] for alternative approaches.

76

Mdadj −CT dadj +KTdadj −GT dadj = fadj, (3.13a)

dadj =

∫ t

0

dadj(τ)|PML dτ, (3.13b)

where superscript “adj” refers to the adjoint problem, dadj = (wTh ,T

Th )

T is the vector

of nodal unknowns comprising discrete values of w in ΩRD∪ΩPML and discrete values

of T only in ΩPML, and fadj is a vector comprising the misfit at receiver locations.

Moreover, system matricesM, C, K, G, are identical to those of the forward problem

and, thus, with minor adjustments, an implementation of the forward problem can

also be used for the solution of the adjoint problem.

The matrixM in (3.13) can be diagonalized by using spectral elements, similar

to what we did in (3.5). We rewrite (3.13) as a first-order system

d

dt

y0

y1

My2

=

0 I 00 0 IGT −KT CT

y0

y1

y2

+

00fadj

, (3.14)

where y0 = dadj, y1 = dadj, y2 = dadj, with final values y0(T ) = 0, y1(T ) = 0, and

y2(T ) = 0. We use an explicit RK-4 method to integrate (3.14) in time. The scheme

is outlined in Appendix B.3.

3.2.1.3 The control problems

Lastly, we take the derivative of L with respect to λ and µ in directions λ

and µ, which yields the reduced gradients with respect to λ and µ, respectively. We

77

restrict the reduced gradients to ΩRD (the material properties at the interfaces ΓI

are extended into the PML buffer). For the TN regularization, this yields

L′(u,S,w,T, λ, µ)(λ) = Rλ

ΩRD

∇λ · ∇λ dΩ−∫ T

0

ΩRD

λ ∇w :(divu)I dΩ dt, (3.15a)

L′(u,S,w,T, λ, µ)(µ) = Rµ

ΩRD

∇µ · ∇µ dΩ−∫ T

0

ΩRD

µ ∇u :[∇w + (∇w)T

]dΩ dt.

(3.15b)

Setting the above derivatives to zero, results in the control problems. Similarly, for

the TV regularization, the control problems read

L′(u,S,w,T, λ, µ)(λ) = Rλ

ΩRD

∇λ · ∇λ(∇λ · ∇λ+ ǫ)

1

2

dΩ−∫ T

0

ΩRD

λ ∇w :(divu)I dΩ dt,

(3.16a)

L′(u,S,w,T, λ, µ)(µ) = Rµ

ΩRD

∇µ · ∇µ(∇µ · ∇µ+ ǫ)

1

2

−∫ T

0

ΩRD

µ ∇u :[∇w + (∇w)T

]dΩ dt. (3.16b)

Discretization of either (3.15) or (3.16) result in the following form

Mgλ = Rλ gλreg + gλ

mis, (3.17a)

Mgµ = Rµ gµreg + gµ

mis, (3.17b)

78

where M is a mass-like matrix, gλ and gµ is the vector of discrete values of the

(reduced) gradient for λ and µ, respectively, and gλreg, g

µreg and gλ

mis, gµmis are the

associated vectors corresponding to the regularization-part and misfit-part of gλ and

gµ. We refer to Appendix A.4 for matrix and vector definitions, and discretization

details.

3.2.2 The inversion process

A solution of (3.6) requires simultaneous satisfaction of the state problem

(3.5), the adjoint problem (3.14), and the control problems (3.17). This approach

–a full-space method– is, in principle, possible [91]; however, the associated com-

putational cost can be substantial. Alternatively, a reduced-space method may

be adopted, in which, discrete material properties are updated iteratively, using

a gradient-based minimization scheme. The latter approach is employed here, and is

discussed next.

We start with an assumed initial spatial distribution of the control parameters

(λ and µ), and solve the state problem (3.5) to obtain dst = (uTh ,S

Th )

T . With the

misfit known, we solve the adjoint problem (3.14) and obtain dadj = (wTh ,T

Th )

T . With

uh and wh known, the (reduced) material gradients gλ and gµ, can be computed from

(3.17). Thus, the vector of material values, at iteration k + 1, can be computed by

using a search direction via

79

λk+1 = λk + αλk sλk , (3.18a)

µk+1 = µk + αµk s

µk , (3.18b)

where λ and µ comprise the vector of discrete values for λ and µ, respectively, αλk ,

αµk are step lengths, and sλk , s

µk are the search directions for λk and µk. Herein,

we use the L-BFGS method to compute the search directions [92]6. Moreover, to

ensure sufficient decrease of the objective functional at each inversion iteration, we

employ an Armijo backtracking line search [92], which is outlined in Algorithm 1.

The inversion process discussed thus far is summarized in Algorithm 2.

Algorithm 1 Backtracking line search.

1: Choose αλ, αµ, c1, ρ ⊲ e.g., αλ = 1, αµ = 1, c1 = 10−4, ρ = 0.52: while J(λk + αλ sλk ,µk + αµ s

µk) ≥ J(λk,µk) + c1(α

λ gλk · sλk + αµ gµ

k · sµk) do

3: αλ ← ραλ

4: αµ ← ραµ

5: end while6: Terminate with αλ

k = αλ, αµk = αµ

We remark that for the reduced-space method, either (3.15) or (3.16) can also

be expressed as

L′(u,S,w,T, λ, µ)(λ) = J′(λ, µ)(λ), (3.19a)

L′(u,S,w,T, λ, µ)(µ) = J′(λ, µ)(µ), (3.19b)

6In the numerical experiments that we perform, we store m = 15 L-BFGS vectors.

80

Algorithm 2 Inversion for Lame parameters.

1: k ← 02: Set initial guess for material property vectors λk, µk

3: Compute J(λk,µk) ⊲ Eq. (3.6)4: Set convergence tolerance tol5: while J(λk,µk) > tol do6: Solve the state problem for dst = (uT

h ,STh )

T ⊲ Eq. (3.5)7: Solve the adjoint problem for dadj = (wT

h ,TTh )

T ⊲ Eq. (3.14)8: Evaluate the discrete reduced gradients gλ

k , gµk ⊲ Eqs. (3.17)

9: Compute search directions sλk , sµk ⊲ L-BFGS

10: Choose step lengths αλk , α

µk ⊲ Algorithm 1

11: Update material property vectors λk, µk ⊲ Eq. (3.18)12: k ← k + 113: end while

where the equality in (3.19) is due to the satisfaction of the state problem. Therefore,

the reduced gradients in (3.17), are, indeed, the gradients of the objective functional

with respect to λ and µ.

3.2.3 Buttressing schemes

Inverse medium problems are notoriously ill-posed. They suffer from solution

multiplicity; that is, material profiles that are very different from each other, and,

potentially non-physical, can become solutions to the misfit minimization problem.

Regularization of the control parameters alleviates the ill-posedness; however, this

alone, may not be adequate when dealing with large-scale complex problems. In

this part, we discuss additional strategies that further assist the inversion process,

outlined in Algorithm 2, in imaging complex profiles.

81

3.2.3.1 Regularization factor selection and continuation

Computation of the (reduced) gradients (3.17) necessitates selection of the

regularization factors Rλ and Rµ. A common strategy is to use Morozov’s discrepancy

principle [93], where a constant value for the regularization factor is used throughout

the inversion process. Here, we discuss a simple and practical approach that was

initially developed for acoustic inversion [56], and, later, was successfully applied to

problems involving elastic inversion [57].

We start by rewriting the discrete control problem (3.17), either for λ or µ,

in the following generic form

Mg = R greg + gmis, (3.20)

where g refers to the vector of discrete values of the (reduced) gradient, either for λ

or µ, greg and gmis are the associated vectors corresponding to the regularization-part

and misfit-part of g, and R is the regularization factor yet to be determined. The

main idea is that the “size” of R greg should be proportional to that of gmis at each

inversion iteration. We define the following unit vectors for the two components of

the gradient vector

nreg =greg

‖greg‖, nmis =

gmis

‖gmis‖, (3.21)

where ‖ · ‖ denotes the Euclidean norm. Equation (3.20) can then be written as

82

Mg = R ‖greg‖ nreg + ‖gmis‖ nmis (3.22a)

= ‖gmis‖(R‖greg‖‖gmis‖

nreg + nmis

)

= ‖gmis‖(℘ nreg + nmis

), (3.22b)

where,

℘ = R‖greg‖‖gmis‖

. (3.22c)

In (3.22b), for the “size” of ℘ nreg to be proportional to nmis throughout the entire

inversion process, one may choose 0 < ℘ ≤ 1. Once a value for ℘ has been decided,

R can be computed via

R = ℘‖gmis‖‖greg‖

, (3.23)

where ℘ can take large values (e.g., 0.5) at early stages of the inversion process

and, thus, narrow down the initial search space. As the inversion evolves, ℘ can be

continuously reduced (e.g., down to 0.3) to allow for profile refinement.

3.2.3.2 Source-frequency continuation

Using loading sources with low-frequency content result in an overall image

of the medium that lacks fine features. To allow for more details, and fine-tune

the profile, one needs to probe with higher frequency content. Thus, the inversion

process can be initiated with a signal having a low-frequency content and, then,

83

the frequency range can be increased progressively as inversion evolves. This can be

achieved by using a set of probing signals, ordered such that each signal has a broader

range of frequencies than the previous ones. The inversion process then begins with

using the first signal. Upon convergence, the profile is used as a starting point for

the second signal, and the process is repeated for all signals.

3.2.3.3 Biased search direction for λ

Simultaneous inversion for both λ and µ is remarkably challenging [61]. As

we demonstrate in Section 3.3.2, the objective functional (3.6) is more sensitive to

µ, than to λ. Consequently, as the inversion evolves, the µ-profile converges faster

than that of λ. In [57], a biasing scheme was proposed to accelerate the convergence

of the λ-profile, such that, at the early stages of inversion, the search direction for λ

is biased according to that of µ.

The main idea is that due to physical considerations, the λ-profile should be,

more or less, similar to the µ-profile. Hence, during the early inversion iterations,

the search direction for λ is biased according to

sλk ← ‖sλk‖(W

sµk

‖sµk‖+ (1−W )

sλk‖sλk‖

), (3.24)

where W is a weight that imposes the biasing amount. We assign full weight (W = 1)

on µ at the first inversion iteration, and reduce it linearly down to zero as iterates

evolve (say at k = 50). After that, we let λ evolve on its own, according to the

original, unbiased search direction.

84

3.3 Numerical experiments

We present numerical experiments7 with increasing complexity to test the pro-

posed inversion scheme. In the first example, we verify the accuracy of the gradients,

computed by using Algorithm 2. Next, we focus on material profile reconstruction

for heterogeneous hosts, using synthetic data at measurement locations. Specifically,

we consider: (a) a medium with smoothly varying material properties along depth,

to study various aspects of the inversion scheme; (b) a horizontally-layered profile

with sharp interfaces; (c) a horizontally-layered profile with an ellipsoidal inclusion,

using highly noisy data; and (d) a layered profile with three inclusions in an attempt

to implicate arbitrary heterogeneity. Throughout, we use Gaussian pulses to probe

the considered domains:

f(t) = e−( t−µ

σ)2 ,

where the parameters that characterize the load are given in Table 3.1; µ is the

mean, σ is the deviation, fmax is the maximal frequency content of the pulse, tend is

the active duration of the Gaussian pulse, and the load has an amplitude of 1 kPa.

The time history of the loads and their corresponding Fourier spectrum are shown

in Figure 3.1.

7We developed a code in Fortran, using PETSc [81] to facilitate parallel implementation.

85

Table 3.1: Characterization of Gaussian pulses.

Name fmax µ σ tend

p20 20 0.11 0.0014 0.20p30 30 0.08 0.0007 0.15p40 40 0.06 0.0004 0.12

0 0.1 0.2 0.3 0.40

0.25

0.5

0.75

1

t (sec)

f

fmax = 20 Hzfmax = 30 Hzfmax = 40 Hz

0 10 20 30 400

0.02

0.04

0.06

0.08

frequency (Hz)

|f(ω

)|

fmax = 20 Hzfmax = 30 Hzfmax = 40 Hz

Figure 3.1: Time history of the Gaussian pulses and their Fourier spectrum.

86

3.3.1 Numerical verification of the material gradients

Accurate computation of the discrete gradients is crucial for the robustness

of Algorithm 2. The gradients of the objective functional with respect to the control

parameters can be computed either by the optimize-then-discretize, or, the discretize-

then-optimize approach [68]. While the discretize-then-optimize method yields the

exact discrete gradients of the discrete objective functional [94], this is not always

the case with the optimize-then-discretize scheme [95].

In this part, through a numerical experiment, we demonstrate that the dis-

crete gradients that we compute via the optimize-then-discretize technique, are accu-

rate, and equal to the discrete gradients of the discrete objective functional. To this

end, we consider a heterogeneous half-space with a smoothly varying material profile

along depth, given in (3.27), and mass density ρ = 2000 kg/m3, which, after trunca-

tion, is reduced to a cubic computational domain of length and width 24 m × 24 m,

and 45 m depth. A 5 m-thick PML is placed at the truncation boundaries, as

shown in Fig. 3.2. The material properties at the interfaces ΓI are extended into

the PML. The interior and PML domains are discretized by quadratic hexahedral

spectral elements of size 1 m (i.e., 27-noded bricks, and quadratic-quadratic pairs of

approximation for displacement and stress components in the PML, and, quadratic

approximations for the material properties), and ∆t = 9 × 10−4 s. Throughout, for

the PML parameters, we choose αo = 5, βo = 400 s 1, and a quadratic profile for

the attenuation functions, i.e., m = 2. To probe the medium, we consider vertical

stress loads with Gaussian pulse temporal signatures (see Table 3.1), applied on the

surface of the domain over a region (−11 m ≤ x, y ≤ 11 m), whereas receivers that

87

collect displacement response um(x, t) are also located in the same region, at every

grid point. To obtain synthetic data at the receiver locations, we use a model with

identical characteristics and dimensions as detailed above, but, with a refined dis-

cretization; i.e., element size of 0.5 m, and ∆t = 4.5× 10−4 s. The total duration of

the simulation is T = 0.5 s. We compare directional finite differences of the discrete

objective functional, with directional gradients obtained from (3.17). We start by

defining the finite difference directional derivatives

dfdh (λ,µ)(λ) :=J(λ+ hλ,µ)− J(λ,µ)

h, (3.25a)

dfdh (λ,µ)(µ) :=J(λ,µ+ hµ)− J(λ,µ)

h, (3.25b)

where λ and µ is the discrete direction vector for λ and µ, respectively. The direc-

tional derivatives obtained via the control problems (3.17) are

dco(λ,µ)(λ) = λ M gλ, (3.26a)

dco(λ,µ)(µ) = µ M gµ. (3.26b)

Next, we verify that (3.25) and (3.26) produce identical values for several

choices that we make for λ and µ, with regularization factors8 Rλ = Rµ = 0. We

consider perturbations λ or µ: the unit vector is zero everywhere except at the

8Zero values are considered since convergence difficulties that may arise stem from the misfitpart of the objective functional, and not from the regularization part. Nevertheless, we have alsosuccessfully verified the accuracy of the regularization component of the gradients.

88

x y

z

Figure 3.2: Problem configuration for the verification of the gradients.

component corresponding to coordinate (x, y, z) where the directional derivatives

are being computed. The derivatives dco and dfdh with respect to either λ or µ, for

points with coordinates (x, y, z), are presented in Table 3.2. The digits to which

dfdh agrees with dco are shown in bold. Since pointwise perturbations result in small

changes in the objective functional, numerical roundoff influences the accuracy of

the finite difference directional derivatives, as it has also been reported in [95].

The agreement between the two derivatives is remarkable, both for cases 1-4,

where the wavefield is well-resolved, and for cases 5 and 6, where only 10 points per

wavelength are used for spatial discretization.

89

Table 3.2: Comparison of the directional derivatives.

Case fmax (x,y,z) Pert. dco dfdhfield h = 10−3 h = 10−4 h = 10−5

1 20 Hz (1,1,0) λ −3.03500e-9 −3.03416e-9 −3.03496e-9 −3.03501e-92 20 Hz (1,1,0) µ −2.78908e-9 −2.78875e-9 −2.78917e-9 −2.78921e-93 20 Hz (1,1,-40) λ −5.14848e-11 −5.14711e-11 −5.14647e-11 −5.14996e-114 20 Hz (1,1,-40) µ +4.97666e-10 +4.97411e-10 +4.97512e-10 +4.97366e-105 40 Hz (1,1,0) λ −1.07645e-9 −1.07623e-9 −1.07652e-9 −1.07656e-96 40 Hz (1,1,0) µ −1.56155e-9 −1.56153e-9 −1.56178e-9 −1.56180e-9

3.3.2 Smoothly varying heterogeneous medium

We consider a heterogeneous half-space with a smoothly varying material

profile along depth, given by

λ(z) = µ(z) = 80 + 0.45 |z|+ 35 exp(− (|z| − 22.5)2

150

)(MPa), (3.27)

and mass density ρ = 2000 kg/m3, which, after truncation, is reduced to a cu-

bic computational domain of length and width 40 m × 40 m, and 45 m depth.

A 6.25 m-thick PML is placed at the truncation boundaries, as illustrated in Fig.

3.3. The target profiles are shown in Fig. 3.4. The material properties at the

interfaces ΓI are extended into the PML. The interior and PML domains are dis-

cretized by quadratic hexahedral spectral elements of size 1.25 m (i.e., 27-noded

bricks, and quadratic-quadratic pairs of approximation for displacement and stress

components in the PML, and, quadratic approximations for the material properties),

and ∆t = 10−3 s. This leads to 3, 578, 136 state unknowns, and 616, 850 material

parameters. To probe the medium, we consider vertical stress loads with Gaussian

90

pulse temporal signatures (see Table 3.1), applied on the surface of the domain over

a region (−17.5 m ≤ x, y ≤ 17.5 m), whereas receivers that collect displacement

response um(x, t) are placed at every grid point, in the same region.

x y

z

Figure 3.3: Problem configuration: material profile reconstruction of a smoothlyvarying medium.

Before attempting simultaneous inversion for the two Lame parameters, we

perform single parameter inversion for a) µ only, assuming λ is a priori known and

fixing it to the target profile; and b) λ only, assuming the distribution of µ is known.

3.3.2.1 Single parameter inversion

First, we assume λ is a priori known, and fix it to the target profile. We start

inverting for µ, with a homogeneous initial guess of 80 MPa, exploiting Tikhonov

regularization for taming ill-posedness and solution multiplicity. We use the Gaussian

91

80 100 120

70 130

(a)

80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ and µ (MPa)

(b)

Figure 3.4: Smoothly varying medium: (a) target λ and µ (MPa); and (b) profile at(x, y) = (0, 0).

pulse p20 with maximal frequency content fmax = 20 Hz (see Table 3.1) for 50

iterations, and, then, switch to p30 with fmax = 30 Hz. After 156 iterations, µ

converges to the target profile, as shown in Fig. 3.5(b). We compare the inverted

cross-sectional profiles of µ with the target profile at three different locations, as

shown in Figs. 3.5(c), 3.5(d), and 3.5(e). The agreement between the two profiles

is excellent. Reduction of the misfit functional with respect to inversion iterations is

shown in Fig. 3.7(a) (almost 7 orders of magnitude).

Next, we fix µ to the target profile, and invert for λ, starting with a homoge-

neous initial guess of 80 MPa. We use the Gaussian pulse p20 for 160 iterations, p30

up to the 300th iteration, and then switch to p40. After 456 iterations, the optimizer

converges to the profile displayed in Fig. 3.6(a). The agreement between the target

profile and the inverted profile is remarkable. We compare the two profiles at three

92

80 100 120

70 130

(a) λ (a priori known)

80 100 120

70 130

(b) µ (inverted)

70 80 90 100 110 120 130−45

−30

−15

0

µ (MPa)

z(m

)

TargetInvertedInitial guess

(c) µ at (x, y) = (0, 0)

70 80 90 100 110 120 130−45

−30

−15

0

µ (MPa)

z(m

)

TargetInvertedInitial guess

(d) µ at (x, y) = (10, 10)

70 80 90 100 110 120 130−45

−30

−15

0

µ (MPa)

z(m

)

TargetInvertedInitial guess

(e) µ at (x, y) = (20, 20)

Figure 3.5: Single-parameter inversion (µ only) for a smoothly varying medium.

93

different cross-sections shown in Figs. 3.6(c), 3.6(d), and 3.6(e): the agreement be-

tween the two profiles is excellent. The misfit history is shown in Fig. 3.7(b); the

optimizer reduced the misfit almost 6 orders of magnitude.

80 100 120

70 130

(a) λ (inverted)

80 100 120

70 130

(b) µ (a priori known)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(c) λ at (x, y) = (0, 0)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(d) λ at (x, y) = (10, 10)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(e) λ at (x, y) = (20, 20)

Figure 3.6: Single-parameter inversion (λ only) for a smoothly varying medium.

We remark that the initial value of the misfit in the first experiment is almost

2 orders of magnitude more than that of the second experiment. This indicates that

the objective functional is not equally sensitive to both control parameters, as it has

94

0 50 100 150 20010

−10

10−8

10−6

10−4

10−2

100

iteration

misfit

fmax = 20 Hz fmax = 30 Hz

(a) reconstruction for µ only

0 100 200 300 400 50010

−10

10−8

10−6

10−4

10−2

iteration

misfit

fmax = 20 Hz fmax = 30 Hz fmax = 40 Hz

(b) reconstruction for λ only

Figure 3.7: Variation of the misfit functional with respect to inversion iterations(single parameter inversion).

also been reported in [57]: the objective functional is more sensitive to µ.

3.3.2.2 Simultaneous inversion

We start with a homogeneous initial guess of 80 MPa for both λ and µ and

attempt simultaneous inversion. The target profiles are shown in Fig. 3.4, and the

inverted profiles are displayed in Figs. 3.8(a) and 3.8(b). We also compare the cross-

sectional values of the target and inverted profiles at three different locations, shown

in Fig. 3.9. Although the inverted µ profile agrees reasonably well with the target

profile, inversion for λ is not satisfactory, and the inverted profile departs from the

target as depth increases.

Due to the unsuccessful inversion of the λ profile in the case of simultaneous

inversion, in the next experiment, we bias the search direction of λ based on that of

µ, at the very early stages of inversion, according to the procedure detailed in Section

3.2.3.3. This leads to the successful reconstruction of the two profiles, as is shown in

95

80 100 120

70 130

(a) λ (inverted)

80 100 120

70 130

(b) µ (inverted)

Figure 3.8: Simultaneous inversion for λ and µ using unbiased search directions.

Figs. 3.10(a) and 3.10(b). In Fig. 3.11, we compare the cross-sectional values of the

target and the inverted profiles. The agreement of the inverted µ profile with the

target is remarkable. Moreover, the inverted λ profile agrees reasonably well with

the target, with some discrepancies in depth. The misfit history is shown in Fig.

3.12(b), where the kink in the misfit curve at the 50th iteration corresponds to the

termination point of the biasing scheme.

We remark that in practical applications, one is more interested in the shear

wave velocity (cs) and compression wave velocity (cp) profiles. Once the Lame pa-

rameters have been determined, the wave velocities can be readily computed via

cs =

õ

ρ, cp =

√λ+ 2µ

ρ. (3.28)

In Fig. 3.13, we compare the compressional wave velocities at three different cross-

96

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(a) λ at (x, y) = (0, 0)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(b) λ at (x, y) = (10, 10)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(c) λ at (x, y) = (20, 20)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(d) µ at (x, y) = (0, 0)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(e) µ at (x, y) = (10, 10)

70 80 90 100 110 120 130−45

−30

−15

0z

(m)

µ (MPa)

TargetInvertedInitial guess

(f) µ at (x, y) = (20, 20)

Figure 3.9: Cross-sectional profiles of λ and µ using unbiased search directions.

97

80 100 120

70 130

(a) λ (inverted)

80 100 120

70 130

(b) µ (inverted)

Figure 3.10: Simultaneous inversion for λ and µ using biased search directions.

sectional locations, where the agreement between the reconstructed cp profile and the

target is remarkable. The shear wave velocity does not depend on λ, and, therefore,

its quality is similar to that of the µ profile.

3.3.3 Layered medium

We consider a 40 m × 40 m × 45 m layered medium, where a 6.25 m-thick

PML is placed at its truncation boundaries. The properties of the medium are

λ(z) = µ(z) =

80 MPa, for − 12 m ≤ z ≤ 0 m,

101.25 MPa, for − 27 m ≤ z < −12 m,

125 MPa, for − 50 m ≤ z < −27 m,

(3.29)

and are shown in Fig. 3.14, with mass density ρ = 2000 kg/m3. The material prop-

erties at the interfaces ΓI are extended into the PML buffer. The interior and PML

domains are discretized by quadratic hexahedral spectral elements of size 1.25 m

98

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(a) λ at (x, y) = (0, 0)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)λ (MPa)

TargetInvertedInitial guess

(b) λ at (x, y) = (10, 10)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(c) λ at (x, y) = (20, 20)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(d) µ at (x, y) = (0, 0)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(e) µ at (x, y) = (10, 10)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(f) µ at (x, y) = (20, 20)

Figure 3.11: Cross-sectional profiles of λ and µ using biased search directions.

0 50 100 150 20010

−8

10−6

10−4

10−2

100

iteration

misfit

fmax = 20 Hzfmax = 30 Hz

fmax = 40 Hz

(a) unbiased search directions

0 100 200 300 400 500 600 70010

−10

10−8

10−6

10−4

10−2

100

iteration

misfit

fmax = 20 Hz fmax = 30 Hz fmax = 40 Hz

(b) biased search direction for λ

Figure 3.12: Variation of the misfit functional with respect to inversion iterations(simultaneous inversion).

99

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(a) cp at (x, y) = (0, 0)

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(b) cp at (x, y) = (10, 10)

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(c) cp at (x, y) = (20, 20)

Figure 3.13: Cross-sectional profiles of cp using biased search directions.

(i.e., 27-noded bricks, and quadratic-quadratic pairs of approximation for displace-

ment and stress components in the PML, and, quadratic approximations for the

material properties), and ∆t = 10−3 s. For probing the medium, we use vertical

stress loads with Gaussian pulse temporal signatures, applied on the surface of the

domain over a region (−17.5m ≤ x, y ≤ 17.5m), whereas receivers that collect dis-

placement response um(x, t) are also located in the same region, at every grid point.

We start the inversion process with a homogeneous initial guess of 80 MPa

for the Lame parameters, and attempt simultaneous inversion for both λ and µ,

using the biasing scheme outlined in Section 3.2.3.3. We use the Total Variation

regularization scheme, with ǫ = 0.01, to capture the sharp interfaces of the target

profiles. We use the Gaussian pulse p20, with fmax = 20 Hz, and final simulation time

T = 0.45 s, for 310 iterations. The resulting profiles are shown in Fig. 3.15(a) and

3.15(b), where the layering of the medium is clearly visible in the inverted profiles.

To improve the quality of the inverted profiles, we use them as an initial guess with

the Gaussian pulse p30, and final simulation time of T = 0.4 s, for up to the 860th

100

80 100 120

70 140

(a)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ and µ (MPa)

(b)

Figure 3.14: Layered medium: (a) target λ and µ (MPa); and (b) profile at (x, y) =(0, 0).

iteration, and, then, switch to p40, with final simulation time of T = 0.4 s. After

1112 iterations, the optimizer converges to the profiles displayed in Fig. 3.15(c) and

3.15(d). There is excellent agreement between the inverted µ profile and the target

profile. The inverted λ profile is also in good agreement with the target profile:

the two top layers have been reconstructed quite well, whereas the bottom layer is

slightly “stiffer” in its middle zone. We compare the inverted profiles with the targets

at three different cross-sections, shown in Fig. 3.16. Due to the TV regularization,

sharp interfaces have been captured quite successfully. In Fig. 3.17, we compare the

cp profile with the target, at the same cross-sections; the agreement between the two

profiles is impressive. Figure 3.18 shows the misfit history: the optimizer reduced

the misfit almost 7 orders of magnitude.

101

80 100 120

70 140

(a) λ (fmax = 10 Hz)

80 100 120

70 140

(b) µ (fmax = 10 Hz)

80 100 120

70 140

(c) λ (fmax = 40 Hz)

80 100 120

70 140

(d) µ (fmax = 40 Hz)

Figure 3.15: Simultaneous inversion for λ and µ (layered medium).

102

70 80 90 100 110 120 130 140 150−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(a) λ at (x, y) = (0, 0)

70 80 90 100 110 120 130 140 150−45

−30

−15

0

z(m

)λ (MPa)

TargetInvertedInitial guess

(b) λ at (x, y) = (10, 10)

70 80 90 100 110 120 130 140 150−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(c) λ at (x, y) = (20, 20)

70 80 90 100 110 120 130 140 150−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(d) µ at (x, y) = (0, 0)

70 80 90 100 110 120 130 140 150−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(e) µ at (x, y) = (10, 10)

70 80 90 100 110 120 130 140 150−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(f) µ at (x, y) = (20, 20)

Figure 3.16: Cross-sectional profiles of λ and µ (layered medium).

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(a) cp at (x, y) = (0, 0)

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(b) cp at (x, y) = (10, 10)

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(c) cp at (x, y) = (20, 20)

Figure 3.17: Cross-sectional profiles of cp (layered medium).

103

0 200 400 600 800 1000 120010

−8

10−6

10−4

10−2

100

iteration

misfit

fmax = 20 Hz fmax = 30 Hz fmax = 40 Hz

Figure 3.18: Variation of the misfit functional with respect to inversion iterations(layered medium).

3.3.4 Layered medium with inclusion

We consider a layered medium with an inclusion. The problem consists of a

40 m × 40 m × 45 m layered medium with an ellipsoidal inclusion, where a 6.25 m-

thick PML is placed at its truncation boundaries. The material profiles are given

by

λ(z) = µ(z) =

80 MPa, for − 12 m ≤ z ≤ 0 m,

101.25 MPa, for − 27 m ≤ z < −12 m,

125 MPa, for − 50 m ≤ z < −27 m,

156.8 MPa, for ellipsoidal inclusion,

(3.30)

and are shown in Fig. 3.19, with constant mass density ρ = 2000 kg/m3. The

ellipsoidal inclusion occupies the region (x−7.57.5

)2 + (y5)2 + ( z+12

5.5)2 ≤ 1. The mate-

rial properties at the interfaces ΓI are extended into the PML buffer. The interior

and PML domains are discretized by quadratic hexahedral spectral elements of size

104

1.25 m, and ∆t = 10−3 s. To illuminate the domain, we use vertical stress loads

with Gaussian pulse temporal signatures, applied on the surface of the medium over

a region (−17.5m ≤ x, y ≤ 17.5m), whereas the receivers are also placed at every

grid point in the same region.

75 100 125 150

60 170

(a)

70 90 110 130 150 170−45

−30

−15

0

z(m

)

λ and µ (MPa)

(b)

Figure 3.19: Layered medium with inclusion: (a) target λ and µ; and (b) profile at(x, y) = (7.5, 0).

We use the Total Variation regularization scheme to alleviate ill-posedness

and solution multiplicity, with ǫ = 0.01. Similar to the previous examples, we use

a source-frequency continuation scheme, starting with the Gaussian pulse p20 with

maximal frequency content of 20 Hz for T = 0.45 s, and, when updates in the ma-

terial profiles become practically insignificant, we switch to the next load in Table

3.1, which contains a broader range of frequencies, and, therefore, is able to image

finer features. Figure 3.20(a) and 3.20(b) show the material profiles after 410 itera-

tions, which adequately capture the layering of the domain as well as the ellipsoidal

105

inclusion. To improve the quality of the reconstructed profiles, we use them as an

initial guess with the Gaussian pulse p30, and final simulation time of T = 0.4 s,

for up to the 730th iteration, and, then, switch to p40, with final simulation time of

T = 0.4 s. Figure 3.20(c) and 3.20(d) show the inverted profiles after 1160 iterations.

The sharp interfaces between the three layers and around the ellipsoidal inclusion

are very well captured for the µ profile. The λ profile agrees reasonably well with

the target, showing some “stiff” features at the center of the bottom layer, similar

to the previous example.

Figures 3.21 and 3.22 compare the inverted profiles with the target profiles

at three different cross-sectional lines of the domain, indicating successful imaging of

both the layering and the inclusion. Variation of the misfit functional with respect

to the inversion iterations is shown in Fig. 3.22, where, again, a kink at the 50th

iteration of the misfit curve, corresponds to the termination point of the biasing

scheme.

Encouraged by the successful performance of the proposed inversion algorithm

with noise-free data, next, we consider adding different levels of Gaussian noise to the

measured synthetic response at the receiver locations, and investigate its effect on the

inversion. Figures 3.24(a)-3.24(d) show the measured displacement response of the

system at (x, y, z) = (3.125, 13.75, 0) m, subjected to the p20 pulse, contaminated

with 1%, 5%, 10%, and 20% Gaussian noise, respectively. Using the source-frequency

continuation scheme, the optimizer converges after 811 and 751 iterations, respec-

tively, for cases corresponding to the 1% and 5% Gaussian noise levels. The inverted

profiles are shown in Fig. 3.25. The reconstruction is successful, with minor dis-

106

75 100 125 150

60 170

(a) λ (fmax = 10 Hz)

75 100 125 150

60 170

(b) µ (fmax = 10 Hz)

75 100 125 150

60 170

(c) λ (fmax = 40 Hz)

75 100 125 150

60 170

(d) µ (fmax = 40 Hz)

Figure 3.20: Simultaneous inversion for λ and µ (layered medium with inclusion).

107

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(a) λ at (x, y) = (7.5, 0)

60 80 100 120 140 160−45

−30

−15

0

z(m

)λ (MPa)

TargetInvertedInitial guess

(b) λ at (x, y) = (7.5, 10)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(c) λ at (x, y) = (20, 20)

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(d) µ at (x, y) = (7.5, 0)

60 80 100 120 140 160−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(e) µ at (x, y) = (7.5, 10)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(f) µ at (x, y) = (20, 20)

Figure 3.21: Cross-sectional profiles of λ and µ (layered medium with inclusion).

320 350 380 410 440 470 500−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(a) cp at (x, y) = (7.5, 0)

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(b) cp at (x, y) = (7.5, 10)

330 350 370 390 410 430 450−45

−30

−15

0

z(m

)

cp (m/s)

TargetInvertedInitial guess

(c) cp at (x, y) = (20, 20)

Figure 3.22: Cross-sectional profiles of cp (layered medium with inclusion).

108

0 200 400 600 800 1000 120010

−8

10−6

10−4

10−2

100

iteration

misfit

fmax = 20 Hz fmax = 30 Hz fmax = 40 Hz

Figure 3.23: Variation of the misfit functional with respect to inversion iterations(layered medium with inclusion).

crepancies on the top surface. Next, we increase the noise level to 10% and 20%,

and attempt inversion; after 770 and 674 iterations, respectively, we converge to the

profiles shown in Fig. 3.26. The quality of the inverted profiles decreases as the noise

level increases. However, similarly to the previous case, except for a thin layer on the

top surface, inversion is successful. In Fig. 3.27, we compare cross-sectional profiles

of λ and µ with the target, at different noise levels, at (x, y) = (7.5, 0) m, which

passes through the center of the ellipsoidal inclusion. Sharp interfaces are captured

remarkably well for the µ profile, even at the presence of 20% noise. The inversion

for λ is also satisfactory.

3.3.5 Layered medium with three inclusions

In the last example, we consider a layered medium, with three inclusions, to

study the performance of our inversion scheme for a more complex material profile.

The problem consists of an 80 m × 80 m × 45 m medium, where a 6.25 m-thick

109

0 0.2 0.4 0.6 0.8 1−0.02

0

0.02

0.04

0.06

0.08

t (sec)

uz

(m)

(a) with 1% Gaussian noise

0 0.2 0.4 0.6 0.8 1−0.02

0

0.02

0.04

0.06

0.08

t (sec)

uz

(m)

(b) with 5% Gaussian noise

0 0.2 0.4 0.6 0.8 1−0.02

0

0.02

0.04

0.06

0.08

t (sec)

uz

(m)

(c) with 10% Gaussian noise

0 0.2 0.4 0.6 0.8 1−0.02

0

0.02

0.04

0.06

0.08

t (sec)

uz

(m)

(d) with 20% Gaussian noise

Figure 3.24: Measured displacement response of the layered medium with inclusion,at (x, y, z) = (3.125,−13.75, 0) m, due to the Gaussian pulse p20, contaminated withGaussian noise.

110

75 100 125 150

60 170

(a) λ (1% Gaussian noise)

75 100 125 150

60 170

(b) µ (1% Gaussian noise)

75 100 125 150

60 170

(c) λ (5% Gaussian noise)

75 100 125 150

Mu

60 170(d) µ (5% Gaussian noise)

Figure 3.25: Simultaneous inversion for λ and µ using measured data contaminatedwith 1% and 5% Gaussian noise (layered medium with inclusion).

111

75 100 125 150

60 170

(a) λ (10% Gaussian noise)

75 100 125 150

60 170

(b) µ (10% Gaussian noise)

75 100 125 150

60 170

(c) λ (20% Gaussian noise)

75 100 125 150

60 170

(d) µ (20% Gaussian noise)

Figure 3.26: Simultaneous inversion for λ and µ using measured data contaminatedwith 10% and 20% Gaussian noise (layered medium with inclusion).

112

60 80 100 120 140 160 180−45

−30

−15

0z

(m)

λ (MPa)

Target1% noise5% noise10% noise20% noise

(a) λ at (x, y) = (7.5, 0)

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

µ (MPa)

Target1% noise5% noise10% noise20% noise

(b) µ at (x, y) = (7.5, 0)

Figure 3.27: Cross-sectional profiles of λ and µ at different noise levels (layeredmedium with inclusion).

PML is placed at its truncation boundaries. The material profiles are given by

λ(z) = µ(z) =

80 MPa, for − 15 m ≤ z ≤ 0 m,

101.25 MPa, for − 30 m ≤ z < −15 m,

125 MPa, for − 50 m ≤ z < −30 m,

156.8 MPa, for spheroidal: (x+203.75

)2 + (y+2020

)2 + ( z+8.753.75

)2 ≤ 1,

156.8 MPa, for ellipsoidal: (x−2015

)2 + (y−207.5

)2 + ( z+305

)2 ≤ 1,

80 MPa, for sphere: (x− 20)2 + (y + 20)2 + (z + 35)2 ≤ 6.25,

and are shown in Fig. 3.28, with constant mass density ρ = 2000 kg/m3. Figures

3.29(a) and 3.29(b) depict the target profiles on a cross-section through the domain

situated at 8.75 m and 35 m from the top surface, going through the ellipsoid’s and

sphere’s midplane, respectively. In terms of the smallest wavelength9 the prescribed

9The smallest wavelength is equal to the smallest velocity in the formation 200 m/s, divided by

113

geometry comprises a domain of 16 × 16 × 9 wavelengths long, wide, and deep, a

spherical inclusion with a diameter equal to 2.5 wavelengths, an ellipsoidal inclusion

of 6 × 3 × 2 wavelengths, and a spheroidal inclusion of 1.5 × 8 × 1.5 wavelengths.

The material properties at the interfaces ΓI are extended into the PML buffer. The

interior and PML domains are discretized by quadratic hexahedral spectral elements

(i.e., 27-noded bricks, and quadratic-quadratic pairs of approximation for displace-

ment and stress components in the PML, and, also, quadratic approximation for

material properties) of size 1.25 m, and ∆t = 10−3 s. This leads to 9, 404, 184 state

unknowns, and 2, 429, 586 material parameters. To illuminate the domain, we use

vertical stress loads with Gaussian pulse temporal signatures, applied on the surface

of the medium over a region (−37.5 m ≤ x, y ≤ 37.5 m), whereas receivers are placed

at every grid point, within the same region as the load.

To narrow the feasibility space and alleviate difficulties with solution mul-

tiplicity, we use the Total Variation regularization, with ǫ = 0.01, combined with

the regularization factor continuation scheme outlined in Section 3.2.3.1, the source-

frequency continuation scheme in Section 3.2.3.2, and the biasing scheme for λ search

directions in Section 3.2.3.3. Specifically, we use the regularization parameter = 0.5

when illuminating the medium with pulse p20 for 60 iterations. Next, we use = 0.4

with pulse p30 up to the 290th iteration. Finally, we use = 0.3 with pulse p40

and stop at the 741st iteration. In all the three cases, the total simulation time is

T = 0.7 s.

the largest probing frequency 40 Hz, i.e., 5 m.

114

75 100 125 150

60 170

(a)

75 100 125 150

60 170

(b)

70 90 110 130 150 170−45

−30

−15

0

z(m

)

λ and µ (MPa)

(c)

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

λ and µ (MPa)

(d)

70 80 90 100 110 120 130−45

−30

−15

0

z(m

)

λ and µ (MPa)

(e)

Figure 3.28: Layered medium with three inclusions: target λ and µ (a) along across-section that cuts through the domain from (x, y) = ( 20, 46.5) to ( 20, 20) to(46.5, 20); (b) along a cross-section that cuts through the medium from (x, y) =(20, 46.5) to (20, 20) to ( 46.5, 20); (c) profile at (x, y) = ( 20, 20); (d) profile at(x, y) = (20, 20); and (e) profile at (x, y) = (20, 20).

115

75 100 125 150

60 170

(a)

80 100 120

70 140

(b)

Figure 3.29: Layered medium with three inclusions: target λ and µ on (a) thez = −8.75 m cross-section; and (b) the z = −35 m cross-section.

Figure 3.30 shows the inverted profile along a cross-section that cuts through

the domain from (x, y) = ( 20, 46.5) to ( 20, 20) to (46.5, 20). The cross section

passes through the larger semi-principal axes of both ellipsoids, and shows very good

reconstruction of the µ profile, and satisfactory inversion of the λ profile. The layering

is sharp, and the ellipsoids are captured well. In Fig. 3.31, a cross section of the

inverted profiles from (x, y) = (20, 46.5) to (20, 20) to ( 46.5, 20) is displayed, where

it passes through the smaller semi-principal axes of the ellipsoids and the center of

the sphere. The ellipsoids are well captured; however, the sphere, which consists of

“soft” materials, can hardly be noticed, especially, in the λ profile. Figure 3.32(a) and

3.32(b) show the inverted profiles on a cross-section through the domain, situated

at 8.75 m from the surface, going through the top ellipsoid’s midplane, and show

satisfactory reconstruction of the ellipsoid. To see the reconstruction of the sphere

116

in more detail, we consider a cross-section, which goes through the sphere’s midplane,

situated at 35 m from the top surface; this is shown in Fig. 3.32(c) and 3.32(d). The

sphere’s footprint is visible in the λ profile, whereas it is more conspicuous in the µ

profile.

75 100 125 150

60 170

(a) λ (fmax = 40 Hz)

75 100 125 150

60 170

(b) µ (fmax = 40 Hz)

Figure 3.30: Simultaneous inversion for λ and µ: cross-section cuts through thedomain from (x, y) = ( 20, 46.5) to ( 20, 20) to (46.5, 20) (layered medium withthree inclusions).

We also compare cross sections of the inverted profiles with the target along

three different lines, which pass through the ellipsoids and the sphere. These are

shown in Fig. 3.33. Overall, the inverted profiles are satisfactory.

3.4 Summary

We discussed a full-waveform-based inversion methodology for imaging the

elastic properties of a soil medium in three-dimensional, arbitrarily heterogeneous,

117

75 100 125 150

60 170

(a) λ (fmax = 40 Hz)

75 100 125 150

60 170

(b) µ (fmax = 40 Hz)

Figure 3.31: Simultaneous inversion for λ and µ: cross-section cuts through thedomain from (x, y) = (20, 46.5) to (20, 20) to ( 46.5, 20) (layered medium withthree inclusions).

semi-infinite domains. The problem typically arises in geotechnical site character-

ization and geophysical explorations, where high-fidelity imaging of the two Lame

parameters (or an equivalent pair) is of interest. Elastic waves are used as probing

agents to interrogate the soil medium, and the response of the medium to these waves

are collected at receivers located on the ground surface. The inversion process relies

on minimizing a misfit between the collected response at receiver locations, and a

computed response based on a trial distribution of the Lame parameters. We used

the apparatus of PDE-constrained optimization to impose the forward wave propaga-

tion equations to the minimization problem, directly in the time-domain. Moreover,

PMLs were used to limit the extent of the computational domain.

To alleviate the ill-posedness, associated with inverse problems, we used reg-

118

75 100 125 150

60 170

(a) λ (fmax = 40 Hz)

75 100 125 150

60 170

(b) µ (fmax = 40 Hz)

80 100 120

70 140

(c) λ (fmax = 40 Hz)

80 100 120

70 140

(d) µ (fmax = 40 Hz)

Figure 3.32: Layered medium with three inclusions: (a) inverted λ profile on thez = 8.75 m cross-section; (b) inverted µ profile on the z = 8.75 m cross-section; (c)inverted λ profile on the z = 35 m cross-section; and (d) inverted µ profile on thez = 35 m cross-section.

119

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(a) λ at (x, y) = (−20,−20)

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(b) λ at (x, y) = (20, 20)

70 90 110 130 150−45

−30

−15

0

z(m

)

λ (MPa)

TargetInvertedInitial guess

(c) λ at (x, y) = (20,−20)

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(d) µ at (x, y) = (−20,−20)

60 80 100 120 140 160 180−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(e) µ at (x, y) = (20, 20)

70 90 110 130 150−45

−30

−15

0

z(m

)

µ (MPa)

TargetInvertedInitial guess

(f) µ at (x, y) = (20,−20)

Figure 3.33: Cross-sectional profiles of λ and µ (layered medium with three inclu-sions).

120

ularization schemes, along with a regularization factor continuation scheme, which

tunes the regularization factor adaptively at each inversion iteration. We discussed

additional strategies to robustify the inversion algorithm: specifically, we used (a)

a source-frequency continuation scheme such that the inversion process evolves by

using low-frequency sources, and, gradually, we use sources with higher frequencies;

and (b) a biasing scheme for the λ-profile, such that, at early iterations of inversion,

the search direction for λ is biased based on that of µ. The latter strategy, in particu-

lar, improves the reconstruction of the material profiles when simultaneous inversion

of the two Lame parameters is exercised. To the best of our knowledge, this is the

first attempt that the two Lame parameters have been successfully reconstructed in

three-dimensional PML-truncated domains.

By comparing directional finite differences of the discrete objective functional,

and directional derivatives obtained via the control problems, we verified the accuracy

and correctness of the material gradients. We reported numerical results demonstrat-

ing successful reconstruction of both Lame parameters for smooth and sharp profiles.

Overall, the framework discussed in this study seems robust, practical, and promis-

ing.

121

Chapter 4

Site characterization using full-waveform inversion

In the preceding chapters, we discussed the theoretical aspects of a full-

waveform-inversion-based methodology for the imaging of the elastic properties of

near-surface deposits. We also addressed the forward problem of numerically simu-

lating the propagation of waves in PML-truncated, arbitrarily heterogeneous, elastic

domains. We presented several PML formulations for the three-dimensional case, nu-

merical results for the forward problem, as well as numerical results for the associated

inverse medium problem using synthetic data.

In this chapter, we focus on the exercising of the developed methodology by

real, as opposed to synthetic, data. There are several applications in engineering

that stand to benefit from a robust full-waveform-inversion methodology: among

these, our focus here is geotechnical site characterization. To this end, we use the

data collected during two field experiments in an attempt to validate the theoretical

development. Specifically, we consider two cases consistent with our code develop-

ment cycle: firstly, we describe the design of a field experiment aimed at reproducing

in the field plane-strain conditions, in order to allow us to exercise two-dimensional

inversion codes. We describe the design, the pre-processing, the field data, and the

inversion results. Moreover, we describe comparisons of the inverted full-waveform

122

based profile against those by using two other imaging methods - the non-invasive

SASW, as well as an invasive CPT test. Secondly, we describe the field experi-

ment and associated inversion results obtained by exercising the three-dimensional

inversion codes.

The remainder of this chapter is organized as follows: first, we review the

mathematical and numerical aspects of the underlying inverse medium problem in

two space dimensions. We remark that there are differences between the two- and

three-dimensional formulation that merit a detailed discussion. The differences stem

from the fact that the number of stretching functions for the PML equals the spatial

dimensionality: this difference, in turn, affects the temporal complexity, and gives

rise to alternative formulations, which could not be trivially obtained by simply re-

ducing the three-dimensional formulations to two dimensions. The two-dimensional

development preceded the research described in this dissertation [57]. However, algo-

rithmic modifications/improvements, the field experiment, and the two-dimensional

inversion results using the field data were all part of the work described herein. Next,

we describe the design, data collection, and full-waveform inversion results based on

a field experiment collected at the UCSB NEES site in Garner Valley, California in

March 2012. Lastly, we conclude with summary remarks.

4.1 The inverse medium problem in two space dimensions

To fix ideas, we refer to the depiction of the driving application shown in

Fig. 4.1: similarly to the three-dimensional case, here too we are interested in recon-

structing the formation’s profile shown in Fig. 4.1(a). Given the arbitrary heterogene-

123

ity of the domain of interest, the problem is inherently three-dimensional. Herein, we

describe the methodology by focusing on the two-dimensional counterpart of the orig-

inal problem, as depicted in Fig. 4.1(b): we accept arbitrary heterogeneity within

a two-dimensional slice, but presume that the properties remain unchanged along

the third dimension, i.e., a plane strain problem. While, the problem, as defined,

departs from the true physical three-dimensional case1, it still contains all the algo-

rithmic and theoretical complexities associated with the three-dimensional problem.

Figure 4.1(c) represents the mathematical idealization of the two-dimensional slice

shown in Fig. 4.1(b): the semi-infinite (two-dimensional) physical domain has been

truncated to a finite one, through the introduction of PMLs.

4.1.1 The forward problem

We use a hybrid approach for the simulation of the wave motion inside a two-

dimensional PML-truncated elastic medium. We refer to [33] and references therein

for the complete development of the method. Here, we repeat only the resulting

coupled system of equations. Accordingly, find u(x, t) in ΩRD ∪ΩPML, and S(x, t) in

ΩPML (see Fig. 4.2 for domain and boundary designations), where u and S reside in

appropriate functional spaces and:

1The problem is still valid in the case of horizontal layers, or even in the case of inclined layers,or even in the case of arbitrary plane heterogeneity.

124

(a) (b)

(c)

Figure 4.1: Problem definition: (a) interrogation of a heterogeneous semi-infinitedomain by an active source; (b) a 2D cross-section of the domain showing one sourceand multiple receivers; and (c) computational model truncated from the semi-infinitemedium via the introduction of PMLs.

125

n

z

Figure 4.2: A PML-truncated semi-infinite domain in two dimensions.

divµ [∇u+ (∇u)T ] + λ(div u) I+ b = ρu, in ΩRD × J,

(4.1a)

divST Λe + ST Λp = ρ ( au+ bu+ cu ), in ΩPML × J,(4.1b)

D [ aS+ bS+ cS ] =1

2[ (∇u)Λe + Λe(∇u)T + (∇u)Λp + Λp(∇u)T ], in ΩPML × J.

(4.1c)

The system is initially at rest, and subject to the following boundary and interface

conditions:

126

µ [∇u+ (∇u)T ] + λ (div u) In = gn, on ΓRDN × J, (4.2a)

ST Λe + ST Λpn = 0, on ΓPMLN × J,

(4.2b)

u = 0, on ΓPMLD × J, (4.2c)

u+ = u−, on ΓI × J, (4.2d)

µ [∇u+ (∇u)T ] + λ(div u) In+ + ST Λe + ST Λpn− = 0, on ΓI × J. (4.2e)

In the above, Λe and Λp are the so-called stretch tensors corresponding to evanescent

and propagating waves, respectively, D[·] is the fourth-order elasticity compliance

tensor, and a, b, c are products of certain elements of the stretch tensors [33, 96].

We remark that, while the three-dimensional PML formulation leads to third-

order in time semi-discrete forms, due to the multiplication of three stretching func-

tions, the two-dimensional formulation maintains the second-order temporal com-

plexity typical of elastodynamics. Moreover, due to the different algebraic manipu-

lation of the constitutive and kinematic equations in (4.1c) when compared to the

approach followed in Chapter 2, the resulting semi-discrete form is symmetric, with

a block tri-diagonal mass matrix.

Following a Galerkin approach, similar to what we described in Section 2.3.1,

the following semi-discrete form results [33]:

M2Dd+C2Dd+K2Dd = f , (4.3)

127

where M2D,C2D,K2D are system matrices and are defined in Appendix A.5, d is

the vector of nodal unknowns comprising displacements in ΩRD ∪ ΩPML, and stress

components only in ΩPML, and f is the vector of applied forces. Several methods can

be used to integrate (4.3) in time; for example, application of the well-known family

of Newmark methods would require for the n+ 1-th time step to solve:

Kdn+1 = fn+1 +M2D(a0dn + a2d

n + a3dn) +C2D(a1d

n + a4dn + a5d

n), (4.4)

where the effective stiffness matrix is

K = a0M2D + a1C2D +K2D, (4.5)

and the velocity-like and acceleration-like updates are given by

dn+1 = a1(dn+1 − dn)− a4d

n − a5dn,

dn+1 = a0(dn+1 − dn)− a2d

n − a3dn, (4.6)

in which a0-a5 are constants whose values depend on the choice of the particular New-

mark method2. Given initial conditions d0 = u0, d0 = v0, use of (4.4-4.6) allows the

integration of the semi-discrete form. Alternatively, (4.4) and (4.6), supplemented

by the initial conditions, can be cast in the following compact form:

Qd = f , (4.7)

2For instance, for a constant average acceleration scheme, a0 = 1/(α ∆t2), a1 = δ/(α ∆t), a2 =1/(α ∆t), a3 = 1/(2 α)− 1, a4 = δ/α− 1, a5 = (δ/2 α− 1)∆t, where δ = 1/2, α = 1/4.

128

where d = [d0 d0 d0 d1 d1 d1 . . .dN dN dN ]T corresponds to the space-time dis-

cretization of the unknown variables and their temporal derivatives (N is the number

of time steps, and di are the spatial degrees of freedom at the i-th time step), and

f = [u0 v0 f0 f1 0 0 . . . fN 0 0]T . The discrete forward operator Q is defined as:

Q =

I 0 0 0 0 0 · · · 0 0 0 0 0 00 I 0 0 0 0 · · · 0 0 0 0 0 0K C M 0 0 0 · · · 0 0 0 0 0 0

L1 L2 L3 K 0 0 · · · 0 0 0 0 0 0a1I a4I a5I −a1I I 0 · · · 0 0 0 0 0 0a0I a2I a3I −a0I 0 I · · · 0 0 0 0 0 0...

......

......

.... . .

......

......

......

0 0 0 0 0 0 · · · L1 L2 L3 K 0 00 0 0 0 0 0 · · · a1I a4I a5I −a1I I 00 0 0 0 0 0 · · · a0I a2I a3I −a0I 0 I

, (4.8)

where

L1 = −a0M2D − a1C2D,

L2 = −a2M2D − a4C2D,

L3 = −a3M2D − a5C2D.

We emphasize that (4.7) is precisely the Newmark algorithm written in a

different form, which is better suited for the solution of the inverse medium problem

at hand, as it will become apparent later. Notice that the first two rows of (4.8)

recover the initial conditions, whereas the third row solves for d0. The fourth row

solves for d1, and the fifth and sixth rows yield d1, d1, respectively. Finally, the last

three rows, solve for dN , and update dN , dN , respectively.

129

4.1.2 The inverse problem

We consider the following PDE-constrained minimization problem:

minλ,µ

J(λ, µ) :=1

2

Nr∑

j=1

∫ T

0

Γm

(u− um)2 δ(x− xj) ds(x) dt+ RTN(λ, µ), (4.9)

subject to the (continuous) forward problem governed by the initial- and boundary-

value problem (4.1a)-(4.2e).

In the above, u is the vertical component of the computed displacement,

and um corresponds to the measured vertical displacement component (obtained via

sensor data processing).

One may use the (formal) Lagrangian approach [86], whereby the forward

problem (4.1a)-(4.2e) is imposed via Lagrange multipliers (adjoint variables) to the

functional (4.9). One then seeks a stationary point to the resulting Lagrangian

functional, by forcing the first-order optimality conditions to vanish. This approach

was discussed in detail in Chapter 3, and is referred to as an optimize-then-discretize

approach, since the optimality conditions are sought first in their continuous form,

followed by spatial and temporal discretization steps3.

Alternatively, one may discretize the continuous constrained optimization

problem (4.9) first, and then compute the corresponding discrete optimality con-

ditions [68, 94]. The procedure is referred to as a discretize-then-optimize approach

3See also [57].

130

[94]. The differences between the two approaches are summarized in the following

diagram:

Jdiscretize

> J

G

optimize∨ discretize

> gotd, gdto,

optimize∨

where J is the discrete objective functional, G is the continuous gradient of the con-

tinuous objective functional J, gotd refers to the discrete gradient obtained through

the optimize-then-discretize approach, and gdto is the discrete gradient computed

through the discretize-then-optimize procedure.

Herein, we opt for the discretize-then-optimize approach for the following rea-

sons: the optimize-then-discretize approach, oftentimes, yields an approximation of

the gradient of the discrete functional (4.9), while the discretize-then-optimize ap-

proach yields the exact gradient of the discrete functional. Although both approaches

involve approximation due to the discretization step, the optimize-then-discretize ap-

proach, sometimes4, does not yield the exact gradient of either the continuous func-

tional, or of the discretized functional [94]. Therefore, the optimize-then-discretize

approach may result in inconsistent gradients, which, in turn, may cause numeri-

cal difficulties; for instance, a downhill direction as determined by the inconsistent

gradient, may actually be an uphill direction of the functional. This may force the

Armijo condition to be violated [92] and, eventually, force the optimizer to stop. The

4The inexact gradient is likely to manifest when the time integrator is “unsymmetric”, as in thecase with the Newmark method [97].

131

discretize-then-optimize approach, however, is more robust and does not suffer from

such problems (see, for example, Chapter 4 in [94] for a comprehensive discussion

and [68] for a systematic treatment and a detailed example).

Discrete optimality conditions

Discretization of the objective functional (4.9) in space and time yields:

minλ,µ

J(λ,µ) :=1

2(d− dm)

T B(d− dm) +Rλ

2λTR λ+

2µTR µ, (4.10)

where d satisfies the discrete forward problem (4.7). Here, λ and µ are discrete

material properties, dm are the discrete space-time measurement data, B is the dis-

cretized (space-time) measurement operator5, and R is the matrix corresponding to

the discretization scheme used for the regularization terms. The discrete Lagrangian

corresponding to (4.10) becomes:

L(d, p,λ,µ) := J(λ,µ)− pT (Qd− f), (4.11)

where p = [r0 q0 p0 r1 q1 p1 . . . rN qN pN ]T is the discrete (space-time) Lagrange

multiplier that enforces the discrete forward problem (Qd = f) as a constraint6. The

discrete optimality conditions for (4.11) are obtained by requiring that the derivatives

of the Lagrangian with respect to each of the variables vanish. Taking the derivative

5B is a block diagonal matrix with ∆t B on the diagonal; ∆t denotes the time step, and B isa square matrix that is zero everywhere except on the diagonals that correspond to a degree offreedom for which measured data are available.

6Though not necessary, qi can be thought of as pi, and ri as pi at the i-th time step.

132

with respect to the Lagrange multiplier p recovers the discrete forward problem (4.7),

which in this context, we refer to as the discrete state equation:

Lp(d, p,λ,µ) = −Qd+ f = 0. (4.12)

We remark that equation (4.12) has the structure discussed earlier in section 4.1.1.

The discrete adjoint equation is obtained by requiring that the derivative of the

discrete Lagrangian with respect to d vanish, that is,

Ld(d, p,λ,µ) = −QT p+B(d− dm) = 0. (4.13)

We remark that (4.13) involves the transpose of Q, and hence, (4.13) is solved by

marching backwards in time. For example, from the last two rows of (4.13), we

obtain the final conditions:

pN = 0, (4.14)

qN = 0, (4.15)

respectively; and the third row from the bottom yields

KT rN = ∆tB(dN − dNm) + a1q

N + a0pN , (4.16)

which can be solved for rN . For time steps n = N,N − 1, · · · , 2, we deduce the

following algorithm:

133

update:

pn−1 = (a3MT2D + a5C

T2D)r

n − a5qn − a3p

n,

qn−1 = (a2MT2D + a4C

T2D)r

n − a4qn − a2p

n. (4.17a)

solve:

KTrn−1 = ∆tB(dn−1 − dn−1m ) + a1q

n−1 + a0pn−1

+ (a0MT2D + a1C

T2D)r

n − a1qn − a0p

n. (4.17b)

Notice that contrary to the conventional application of the Newmark’s method, here,

we first update, and then solve. Finally, the first three rows of (4.13) result in the

following equations:

solve:

MT2Dp

0 = (a3MT2D + a5C

T2D)r

1 − a5q1 − a3p

1 (4.18a)

update:

q0 = −CT2Dp

0 + (a2MT2D + a4C

T2D)r

1 − a4q1 − a2p

1,

r0 = −KT2Dp

0 + (a0MT2D + a1C

T2D)r

1 − a1q1 − a0p

1 +∆tB(d0 − d0m). (4.18b)

The third discrete optimality condition is obtained by setting the derivative of (4.11)

with respect to λ and µ to zero. That is:

134

Lλ(d, p,λ,µ) = Rλ R λ− ∂

∂λ(pT Q d) = 0, (4.19)

Lµ(d, p,λ,µ) = Rµ R µ− ∂

∂µ(pT Q d) = 0, (4.20)

where the terms ∂

∂λ(pT Q d) and ∂

∂µ(pT Q d) can be computed in a straightforward

manner, as outlined in Appendix D.

Next, an iterative procedure may be used such that the discrete material prop-

erties are updated according to a gradient-based scheme. Here, we use a Fletcher-

Reeves conjugate gradient scheme [57], endowed with an Armijo backtracking line

search, to iteratively update the material profiles.

4.2 The field experiment - design considerations

In this section, we discuss the design of a field experiment that will provide

us with field-measured data, which can then be used to exercise the two-dimensional

inversion codes we developed/modified based on the preceding theory. To this end,

we attempt to generate plane strain conditions in the field, so that we then attempt

to invert for the properties of a two-dimensional site slice (Fig. 4.3c). Since the

loads we can impart on the ground surface are really three-dimensional, plane strain

conditions would require loading along densely populated lines (Fig. 4.3a) to emulate

theoretical line loads. In this section, we discuss how this can be accomplished in

a practical manner. Naturally, to replicate plane strain conditions in the field, the

loading is only one of the difficulties: a key assumption we make here is that there

is lateral material homogeneity, that is, the site slice may be, in plane, arbitrarily

135

heterogeneous, but the properties do not change along the direction perpendicular to

the slice. While restrictive, the assumption is realistic for layered sites, is a significant

improvement upon pervasive one-dimensional assumptions of other methodologies,

and, due to prior characterizations, it was a reasonable assumption to make for the

specific field experiment site (Hornsby Bend in Austin, TX).

With these assumptions in mind, the question becomes: how can a line load

be approximated by a sequence of loads in the field, which may resemble point loads,

and are suitably spaced along straight lines whose extent remains finite (line loads are

infinite in extent)? The question is depicted in Fig. 4.3. Therefore, we are interested

in arriving at an estimate of the spacing between the loads, and an estimate of the

finite extent of the line of loads. We focus next on these two questions of spatial load

distribution, but, in parallel, we also address the temporal variability of the load

signal. We sketch the process on the basis of a homogeneous halfspace by drawing

on classical solutions; site-specific conditions will also have an effect on load spacing

and load extent, but when the design is based on minimum expected wavelength,

then the design is conservative.

Our methodology is based on the time-domain point-load analytical solutions

in two-dimensional and three-dimensional space, that is, Lamb’s problem and the

Pekeris-Mooney’s problem, respectively [98]. We use these solutions to derive the

response of a two-dimensional or three-dimensional soil domain due to (temporally)

arbitrary loads. In this way, the problem of designing the experiment reduces to a

parametric study. We study first the effect of truncating the line load from extending

to infinity by finding a suitable truncation length. Then, we replace the truncated

136

(a) 3D heterogeneous halfspace sub-jected to a series of line loads ex-tending to infinity; heterogeneity isin plane only.

Ls

s

s

(b) 3D heterogeneous halfspace sub-jected to a finite number of pointsources spaced apart by s, arrangedalong lines of L total length.

(c) Equivalent 2D halfplane subjected toconcentrated sources.

Figure 4.3: Approximation of a 3D halfspace problem by a 2D halfplane problem.

137

load line with equivalent point sources and determine the appropriate spacing be-

tween them. Fig. 4.3 displays schematically this objective. Finally, we comment on

designing signals that are appropriate for probing geotechnical sites and also discuss

theoretical and practical issues that arise in such field experiments.

4.2.1 Line load truncation and spacing requirements

An impulsive in-plane point load is applied on the surface of a halfplane and

the resulting displacements are sought. This plane strain problem is equivalent to

an infinite line load applied on the surface of a halfspace. We denote the vertical

displacement of a point (x1, 0) at time t by G2D(x1, t) , where G2D stands for the

Green’s function, and the impulse load acts at the origin. The temporal dependence

of the load can be described via the Dirac-delta function: exploiting superposition,

we obtain the response due to any arbitrary time signal, denoted by f 2D(t), via the

following convolution integral:

u2D(x1, t) =

∫ t

0

f 2D(τ) G2D(x1, t− τ) dτ. (4.21)

The above relation allows us to compute the response of a halfplane subjected to

any temporally arbitrary vertical force that acts on its surface. Special care must be

taken when computing the integral since it has a singularity due to the arrival of the

Rayleigh wave. Therefore, the integral should be interpreted in the Cauchy principal

value sense. Details of the Green’s function G2D(x1, t) can be found in [98].

A vertical point source that varies as a step function in time is applied on

the surface of a halfspace. The vertical displacement of a point (x1, x2, 0) at time t

138

is denoted by G3D(x1, x2, t), and the load acts at the origin. Details of the Green’s

function G3D are given in Appendix F. The time signal can be described via a

Heaviside function. However, we are interested in obtaining the response due to

any arbitrary load, not only those with a simple step-like signature. Hence, our

first attempt in making this problem fit into our needs, is to represent any arbitrary

load f 3D(t) as a summation of Heaviside functions and denote this approximation

by f 3Dn (t). Indeed, given n+ 1 pairs ti, fi = f 3D(ti), i = 0, 1, · · · , n, we have

f 3Dn (t) =

1

2(f0 + f1)(H0 −H1) +

1

2(f1 + f2)(H1 −H2) + · · ·

+1

2(fn−1 + fn)(Hn−1 −Hn), (4.22)

where we used Hi ≡ H(t− ti) for notational simplicity, and ti indicates time value at

node i. Re-arranging the above relation yields the following more convenient form:

f 3Dn (t) =

n∑

i=0

hi H(t− ti), (4.23)

where

hi =

12(f0 + f1), for i = 0,−1

2(fn−1 + fn), for i = n,

12(fi+1 − fi−1), otherwise.

(4.24)

Next, we consider a uniform distribution of point sources, with a Heaviside

time signature, along a line of finite length. Without loss of generality, we assume

that the sources are positioned symmetrically about the origin, and occupy a total

139

length of 2L along the x2 axis (see Fig. 4.4). The response of the halfspace to this

distribution, at any arbitrary point on the surface along the x1 axis, at time t, is

denoted by G2L(x1, t), and is obtained by the following relation

x1

x2

x3

2L

Figure 4.4: Line load with a finite length.

G2L(x1, t) =

∫ L

−L

G3D(x1, x2, t) dx2, (4.25)

where G2L(x1, t) may be interpreted as the Green’s function of a truncated line

load. We expect that, at the limit as L → ∞, G2L reduces to the solution of the

corresponding plane strain problem in the x1 − x3 plane. Analytical expressions for

the integral exist, and are discussed in detail in Appendix F. Finally, combining

(4.23) and (4.25), we obtain the response of the halfspace due to any arbitrary in

time, but uniform in space, load with a total length of 2L as:

u2L(x1, t) =

n∑

i=0

hi G2L(x1, t− ti). (4.26)

140

Relation (4.26) can be compared against (4.21) to find the appropriate truncation

length L.

Once L is determined, the next step is to replace the continuous line load,

with point sources. This is readily available by combining G3D(x1, x2, t) with (4.23):

if we consider 2m+1 point sources, symmetrically positioned along the x2 axis, and

spaced s distance apart such that s = L/m, the response of the halfspace at a point

along the x1 axis, may be obtained via:

u2Ls (x1, t) =

n∑

i=0

s hi G3D(x1, 0, t− ti) + 2

m∑

j=1

n∑

i=0

s hi G3D(x1, js, t− ti). (4.27)

Relations (4.21) and (4.27) can be compared against each other to determine

the appropriate spacing between point sources.

4.2.2 Verification

We consider two numerical experiments to verify the derivation and numerical

implementation of (4.21) and (4.26). We consider the material properties summa-

rized in Table 4.1 for the homogeneous, isotropic elastic medium under consideration

(notice not all of these properties are independent).

P-wave velocity cp = 346.4 m/sS-wave velocity cs = 200.0 m/sR-wave velocity cR = 183.9 m/sPoisson’s ratio ν = 0.25

Table 4.1: Material properties used in load verification examples.

Lamb and Pekeris-Mooney’s Green’s functions are displayed in Fig. 4.5 where

the observer is located at x1 = 20 m away from the source. The P-wave arrives

141

first, followed by the arrival of the S-wave with a change in slope. The Rayleigh

surface wave comes next, and results in an infinite displacement that correspond

to the singularity of the associated Green’s function. Moreover, in the case of the

Pekeris-Mooney’s problem, the steady-state response that follows after the Rayleigh

singularity corresponds to the Boussinesq’s solution to a static load, acting on a

halfspace.

0 0.1 0.2 0.3 0.4−4

−2

0

2

4x 10

−7

t (sec)

uz

(m)

(a) Lamb’s solution

0 0.1 0.2 0.3 0.4−1

−0.5

0

0.5

1x 10

−10

t (sec)

uz

(m)

(b) Pekeris-Mooney’s solution

Figure 4.5: 2D and 3D Green’s functions.

Example 1. We consider a suddenly applied vertical load, with f(t) = H(t).

Then, in (4.21) we have f 2D(t) = 1, and only one term of the series in (4.26) is

sufficient, namely h0 = 1, while t0 = 0.

Considering the value of the P-wave velocity, if we desire to plot the response

up to 0.4 seconds, the farthest point from the observer that contributes to the re-

sponse is located 140 m away. This means that taking L = 140 m should yield

142

identical results when comparing (4.21) and (4.26). Indeed, this is the case as shown

in Fig. 4.6. We also include the case of L = 50 m, which clearly deviates from the

exact and the L = 140 m case. The deviation begins at t ≈ 0.14 s, and is more

pronounced at around t = 0.27 s. These times correspond to the arrival of the P

and S-wave from the farthest loaded point (i.e. x2 = 50 m), respectively. Finally,

the response reached a steady state value at around 0.29 s, which corresponds to the

arrival of the Rayleigh wave from the farthest point. Therefore, we conclude that

for this example, the discrepancy can mostly be attributed to the arrival of Rayleigh

waves from the farthest point, followed by the final arrival of the S-wave, whereas

the P-wave has negligible effect in this regard.

0 0.1 0.2 0.3 0.4−3

0

3

6x 10

−9

t (sec)

uz

(m)

LambPekeris, L = 140Pekeris, L = 50

Figure 4.6: Comparison of 2D and 3D systems due to a suddenly applied load.

Example 2. We consider a rectangular pulse such that f(t) = H(t)−H(t−

0.2). Hence, two terms of the series in (4.26) are sufficient, i.e., h0 = 1, and h1 = −1,

143

while t0 = 0, and t1 = 0.2. We compare (4.21) and (4.26) for the same L values and

observer location as in the previous example.

0 0.1 0.2 0.3 0.4−3

0

3

6x 10

−9

t (sec)

uz

(m)

LambPekeris, L = 140Pekeris, L = 50

Figure 4.7: Comparison of 2D and 3D systems due to a rectangular pulse load.

Results are displayed in Fig. 4.7. The agreement is excellent between the

two relations when L is sufficiently large; discrepancies result from choosing small

values for L. We emphasize that the two singularities in the plot correspond to the

particular character of the load which initiates and terminates sharply.

4.2.3 Signal design

In this section, we comment on the main features that must be considered

in designing signals for site characterization applications based on full waveform-

based inversion. For the field experiment we conducted we used the Vibroseis trucks

of the NEES@UTexas site (National Science Foundation, Network for Earthquake

144

Engineering Simulation). In particular, we used T-Rex (tri-axial Vibroseis), and

Liquidator (low-frequency Vibroseis) (Fig. 4.8).

Figure 4.8: NEES@UTexas Liquidator Vibroseis.

T-Rex can apply vertical loads with a maximum force amplitude of 267 kN within a

frequency range from 12 Hz to about 180 Hz. It can also be used for applying loads

outside this frequency range with a lower force amplitude. Liquidator is, however,

more widely used when low frequency loading is desired. Liquidator is capable of

applying loads within a frequency range of 1.3 Hz to 75 Hz, with a peak force

amplitude of 89 kN. Thus, from a practical standpoint, the desired frequency content

and amplitude of loading should be restricted within the aforementioned ranges. To

record the motion on the ground surface we used 1 Hz geophones.

Our forward and inverse simulators are based on finite element discretization

of the geotechnical site of interest. High frequency probing waves require a fine mesh

resolution, and, thus, result in an increased computational cost. Therefore, they

should be avoided whenever possible, unless fine features of the formation are of

145

interest. Signals with long time duration require a longer observation period, thus

also resulting in increased computation time, and increased storage requirements

(the solution history must be stored at every time step for inverse problems [94]).

Therefore, the time duration of a signal should be only long enough to effectively

probe the depth of interest.

We favor signals that probe the geotechnical site of interest more effectively.

Typically, these are signals that encompass a range of frequencies rather than con-

taining only few isolated frequencies. The most commonly used class of these signals

are chirps, and have shown their effectiveness in radar and geophysical applications.

In this experiment, we use linear chirps of the form:

f(t) = sin(2π (f0 +kt

2) t), (4.28)

where f0 is the starting frequency, and k is the chirp rate. With these two parameters,

we can design a signal that has a desired frequency range. The starting frequency

may be limited according to the geophone’s resonant frequency, and k is determined

according to the upper bound of the desired frequency range and total time duration

of the signal. We consider four different chirp-type signals with an active and total

time duration of 5 s and 8 s, respectively, which in total, span a frequency range of

3 Hz to 35 Hz. These are summarized in Table 4.2.

The dominant frequencies of the chirp lie between fmin and fmax. For example,

Fig. 4.9 shows chirp’s C-3-8 time dependence and frequency spectrum, where the

strong components are clearly concentrated in the 3 Hz to 8 Hz range.

146

Chirp name f0 k fmin fmax

C-3-8 3 1 3 8C-8-20 8 2.4 8 20C-20-25 20 1 20 25C-25-35 25 2 25 35

Table 4.2: Chirp signals used in the field experiment.

0 1 2 3 4 5 6−2

−1

0

1

2

t (sec)

f

(a) Signal time-history

0 5 10 15 200

0.02

0.04

0.06

0.08

frequency (Hz)

|f(ω

)|

(b) Fourier spectrum

Figure 4.9: Chirp with dominant frequencies between 3 Hz and 8 Hz.

147

Finally, we remark that the signals in Table 4.2 cannot be applied on the

ground surface by either T-rex or Liquidator in their exact form. Indeed, the equip-

ment is only able to exert a load which is “close” to the design load. Therefore,

it becomes necessary to measure the exact applied load due to its significance in

the inversion process. This is done by installing accelerometers on the baseplate

and reaction mass of the Vibroseis equipment. The actual applied load can be ob-

tained by adding the products of the baseplate mass and the reaction mass by their

corresponding measured accelerations.

4.2.4 Parametric studies

In the preceding sections, we developed all the tools we need for the parametric

studies that we carry out in this section. We use (4.23) to represent signals considered

in Table 4.2 with their corresponding Heaviside expansion. We then use (4.26) with

a sufficiently large value for L to obtain the response of a 2D system and use it

as benchmark to find smaller values for L that yield comparable results. Once an

appropriate value for L is determined, (4.27) can be used to find out a suitable

spacing between equivalent point sources.

In order to determine the appropriate length for the line load, we consider

three representative values for L: 50 m, 100 m, and 150 m. We compare the response

of these cases against the benchmark solution (large L value) due to the four loads

considered in Table 4.2. This leads to twelve numerical experiments which provide

insight on how the frequency content of the load may play a role in selecting L.

Results for the first two chirps of Table 4.2 are shown in Fig. 4.10. Each

148

plot depicts two curves, one obtained using the infinite line load length, and the

other obtained using a finite value for L. The agreement between the two curves

is remarkable for all cases, considering that the observation period is 6 s, and the

line load has a relatively short length. We note that for an exact match up to 6 s,

a line load of length L = 2100 m is required. We also observe that larger values of

L yield better results, which is indeed intuitive. Results of the last two chirps of

Table 4.2 also follow a similar trend. We choose the case of L = 100 m for further

investigations in the subsequent part.

We replace next the continuous line load with equivalent, equidistant point

sources, spaced s apart. In particular, we consider two values for s, 5 m, and 10 m,

respectively, and run numerical experiments with the chirp signals of Table 4.2. The

results for the first two chirps of Table 4.2 are shown in Fig. 4.11 and demonstrate

good agreement between the case of a line load of infinite length and that of point

sources, positioned 5 meters apart from each other. The agreement is better for lower

frequencies and deteriorates for higher frequencies. Hence, we choose L = 100 m with

s = 5 m for the first two chirp signals of Table 4.2 in our field experiment (the 10 m

spacing would lead to inaccuracies).

4.2.5 The experiment layout

We discuss next the actual field experiment aimed at a local characterization

of the Hornsby Bend site located in Austin, Texas, using the field data, and the

inversion methodology code discussed. As discussed already, a key assumption is

that the site enjoys symmetry along the sensor plane, as displayed in Fig. 4.3. Our

149

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 50

(a) L=50, L=∞ for Chirp C-3-8

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 100

(b) L=100, L=∞ for Chirp C-3-8

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 150

(c) L=150, L=∞ for Chirp C-3-8

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 50

(d) L=50, L=∞ for Chirp C-8-20

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 100

(e) L=100, L=∞ for Chirp C-8-20

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 150

(f) L=150, L=∞ for Chirp C-8-20

Figure 4.10: Line load truncation effect for different chirps.

150

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 100, s = 5

(a) s =5, chirp C-3-8

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 100, s = 10

(b) s =10, chirp C-3-8

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 100, s = 5

(c) s =5, chirp C-8-20

0 1 2 3 4 5 6−4

−2

0

2

4x 10

−9

t (sec)

uz

(m)

L = ∞L = 100, s = 10

(d) s =10, chirp C-8-20

Figure 4.11: Comparison of infinite line load (L = ∞), a continuous line load offinite length (L = 100 m), and a series of point loads spaced s meters apart over adistance of 100 m.

151

objective is to explore a site of length 200 m along this plane. We place sensors every

5 m. These are shown in Fig. 4.12 with bullets. We consider sources, also spaced

5 m apart from each other, indicated by squares, which are placed along five lines of

length 100 m. Due to the symmetry assumptions, we consider sources only on one

side of the sensor plane, and assume that if we had sources on the mirror side, they

would have yielded the same response as their existing counterparts. The experiment

was performed in Austin, TX, Hornsby Bend in October 2010 (Fig. 4.13).

source locationsensor location

x1,ξ 1

x2,ξ 2

(x1,x 2) (ξ1,ξ 2)

10

0 m

200 m

5 m

5 m

Figure 4.12: The field experiment layout.

4.3 Field experiment records and data processing

In this section, we discuss a subset of the field recorded data, and outline the

data processing procedure. The latter consists of two main parts. First, identifica-

tion and reduction of noise effects associated with the recorded data, and, secondly,

integrating the data in such a way that they readily fit the requirements of our

152

(a) (b)

Figure 4.13: Hornsby Bend field experiment: (a) instrumentation van; and (b) T-rexat the site.

two-dimensional inversion codes.

4.3.1 Signal processing

Our records are inevitably contaminated by noise, which may distort the sig-

nals both at high and low frequencies. Our aim is to identify the parts of a signal

(in the frequency domain), where the signal to noise ratio is low, and filter out these

parts from the records. This is particularly important when dealing with low fre-

quency noise, since the noise gets amplified when the velocity record is integrated

to yield displacement time-history. The latter, may look unphysical and necessitate

baseline adjustment [99], which will have serious consequences in the inversion pro-

cess. On the other hand, high frequency noise may be less of a concern due to the

regularization terms in the inversion algorithm: the regularization terms make the

objective functional less sensitive to the high frequency noise [93].

153

A sampling frequency of 820 Hz is used for digital data collection. Assuming

the highly unlikely event that data being contaminated by noise up to a range of

100 Hz (noise was observed up to 70 Hz), a sampling rate of 200 Hz could prevent

aliasing effects according to the Nyquist sampling theorem [100]. In this sense, we

oversampled the data, which causes no harm. Moreover, in order to reduce the effects

of ambient noise, we repeated each loading five times, and use the average in our

analysis.

We favor finite-impulse-response (FIR) filters since they preserve a signal’s

phase information (linear phase), and do not result in phase distortion, as commonly

occurs in more popular infinite-impulse-response (IIR) filters [100]. We use Matlab’s

equiripple bandpass filter, with high and low cuts of 2.5 Hz and 12 Hz, respectively,

and high and low slopes of 60 dB/octave and 38 db/octave, for the C-3-8 chirp (see

Table 4.2). For the C-8-20 chirp, we use the same type of filter, with high and low

cuts of 2.5 Hz and 25 Hz, and high and low slopes of 60 dB/octave and 26 db/octave,

respectively.

Next, we present some of the field experiment records both in their unpro-

cessed and processed form. For example, Fig. 4.14 shows the C-3-8 chirp, applied

by Liquidator at (ξ1, ξ2) = (0, 0). The record may be compared with the record

in Fig. 4.9, which is the corresponding theoretical curve. Notice that the applied

load lies mainly within the design frequency range, except for the relatively small-

component high-frequency noise, which probably originates with the Liquidator’s

engine and hydraulics.

Figures 4.15, 4.16, and 4.17 depict a subset of the recorded sensor data: both

154

−1 0 1 2 3 4 5 6 7−8

−4

0

4

8x 10

4

t (sec)

f(N

)

(a) Time-history (recorded)

0 5 10 15 20 25 300

1000

2000

3000

4000

5000

frequency (Hz)

|f(ω

)|

(b) Fourier spectrum (recorded)

−1 0 1 2 3 4 5 6 7−8

−4

0

4

8x 10

4

t (sec)

f(N

)

(c) Time-history (filtered)

0 5 10 15 20 25 300

1000

2000

3000

4000

5000

frequency (Hz)

|f(ω

)|

(d) Fourier spectrum (filtered)

Figure 4.14: Force (chirp C-3-8) applied by Liquidator at (ξ1, ξ2) = (0, 0).

unprocessed and processed (filtered) velocity records are shown for various sensor

locations due to different loads. In all cases shown, the load is the chirp C-3-8, as

shown earlier in Fig. 4.14. Shown in Figs. 4.15-4.17 are the velocity time-histories at

(−5, 0, 0), while the load is applied at (0, 0, 0), (0,−5, 0), and (0,−10, 0), respectively.

Geometric decay is noticeable and amplitude reduction in velocity time-history is

observed as distance between the source and observer increases.

4.3.2 Data integration

In this section, we address how we use the, essentially, three-dimensional field

data, in order to obtain records suitable for exercising our two-dimensional codes.

We follow the same lines as in the experiment design. Similarly to what we did in

155

−1 0 1 2 3 4 5 6 7−2

−1

0

1

2x 10

−3

t (sec)

v(m

/s)

(a) Time-history (recorded)

0 5 10 15 20 25 300

0.25

0.5

0.75

1x 10

−4

frequency (Hz)

|f(ω

)|

(b) Fourier spectrum (recorded)

−1 0 1 2 3 4 5 6 7−2

−1

0

1

2x 10

−3

t (sec)

v(m

/s)

(c) Time-history (filtered)

0 5 10 15 20 25 300

0.25

0.5

0.75

1x 10

−4

frequency (Hz)

|f(ω

)|

(d) Fourier spectrum (filtered)

Figure 4.15: Velocity (due to force C-3-8 at (0, 0, 0)) measured at (−5, 0, 0).

(4.27), we obtain the following equivalent two-dimensional velocity time history v2L

from the three-dimensional field-recorded v3D := v3D(x1, x2, t; ξ1, ξ2):

v2L(x1, t; ξ1, ξ2) = v3D(x1, 0, t; ξ1, 0) + 2m∑

j=1

v3D(x1, 0, t; ξ1, js), (4.29)

where x1 denotes a geophone’s location along the x1 axis, (ξ1, ξ2) denotes the load

location, m (=20) is the number of source locations for which x2 > 0, and s = 5 m

is the distance between the loads. v2L is the equivalent two-dimensional velocity

record, which can then be integrated in time to yield the displacement time history.

Similarly, for the equivalent two-dimensional force time history, we obtain:

156

−1 0 1 2 3 4 5 6 7−2

−1

0

1

2x 10

−3

t (sec)

v(m

/s)

(a) Time-history (recorded)

0 5 10 15 20 25 300

0.25

0.5

0.75

1x 10

−4

frequency (Hz)

|f(ω

)|

(b) Fourier spectrum (recorded)

−1 0 1 2 3 4 5 6 7−2

−1

0

1

2x 10

−3

t (sec)

v(m

/s)

(c) Time-history (filtered)

0 5 10 15 20 25 300

0.25

0.5

0.75

1x 10

−4

frequency (Hz)

|f(ω

)|

(d) Fourier spectrum (filtered)

Figure 4.16: Velocity (due to force C-3-8 at (0,−5, 0)) measured at (−5, 0, 0).

f 2L(ξ1, t) =1

(2m+ 1)s

[f 3D(ξ1, 0, t) + 2

m∑

j=1

f 3D(ξ1, js, t)

], (4.30)

where f 3D(ξ1, ξ2, t) denotes the measured force applied at any given location (ξ1, ξ2).

For example, the equivalent line load corresponding to chirp C-3-8, applied at ξ1 = 0,

and the resulting velocity time history at x1 = −5 m are depicted in Fig. 4.18. The

data, both measured force and recorded response, with the aid of (4.29) and (4.30),

can now be readily used for inversion.

157

−1 0 1 2 3 4 5 6 7−2

−1

0

1

2x 10

−3

t (sec)

v(m

/s)

(a) Time-history (recorded)

0 5 10 15 20 25 300

0.25

0.5

0.75

1x 10

−4

frequency (Hz)

|f(ω

)|

(b) Fourier spectrum (recorded)

−1 0 1 2 3 4 5 6 7−2

−1

0

1

2x 10

−3

t (sec)

v(m

/s)

(c) Time-history (filtered)

0 5 10 15 20 25 300

0.25

0.5

0.75

1x 10

−4

frequency (Hz)

|f(ω

)|

(d) Fourier spectrum (filtered)

Figure 4.17: Velocity (due to force C-3-8 at (0,−10, 0)) measured at (−5, 0, 0).

−1 0 1 2 3 4 5 6 7−2

−1

0

1

2x 10

4

t (sec)

f(N

/m

)

(a) Force time-history

−1 0 1 2 3 4 5 6 7−4

−2

0

2

4x 10

−3

t (sec)

v(m

/s)

(b) Velocity time-history

Figure 4.18: Equivalent line load (chirp C-3-8) applied at ξ1 = 0 m, and correspond-ing response at x1 = −5 m.

158

4.4 Inversion results using field experiment data

In this section, we use the inversion theory discussed earlier, along with mea-

sured data from the field experiment, to arrive at an estimate of the spatial distri-

bution of the P and S wave velocities at the target site (Hornsby Bend).

The target domain is a two-dimensional slice, 200 m wide and 48 m deep.

The domain is surrounded on its sides and bottom by a 10 m-thick PML to absorb

outgoing waves. We use bilinear quadrilateral elements with element size of 1 m

when applying the C-3-8 chirp (see Table 4.2), whereas 8-noded serendipity elements

of the same size are used for higher frequency chirps. In all cases, we use 1 m × 1 m

bilinear quadrilateral elements to interpolate the material properties.

According to the experiment layout in Fig. 4.12, we apply loads at every 5 m

along five lines. T-Rex was used for loading along the lines x1 = −60 m, +30 m,

and +60 m, whereas Liquidator was used to load along x1 = −30 m, and 0 m. We

used 36 geophones with a resonant frequency of 1 Hz at every 5 m, along the x1 axis.

After processing the recorded data, per the discussion of the preceding section, we

proceeded with the inversion process.

4.4.1 Inversion process

The inversion process begins with an initial profile for both of the Lame

parameters (linear in depth or homogeneous), and iteratively updates the profile until

the misfit between the measured response and the computed response obtained at

each inversion iteration is minimized. The convergence rate of the inversion process

to the target profile, and even the success of the process itself, depends greatly

159

on the initial guess, as is typically the case. An initial profile, which is close to the

target profile, will likely need fewer number of iterations to converge, compared to an

arbitrary initial profile. This fact can be exploited to speed up the convergence. If, for

example, during a field experiment, an SASW experiment is performed in addition

to the data collection for the full waveform inversion approach, then the SASW-

rendered profile could be used as initial guess for the full waveform inversion. The

SASW profile will be, by definition, horizontally layered, whereas the full waveform-

based inverted profile will be, in general, arbitrarily heterogeneous.

We start the inversion process by applying the first (measured) equivalent

force corresponding to chirp C-3-8 (see Table 4.2 for the theoretical curve and

Fig. 4.18 for the actual, equivalent measured force). There are 5 loads at our dis-

posal, and 36 measuring locations for every load. Owing to linearity, we apply all 5

loads simultaneously and add their corresponding responses at every sensor location.

Other possibilities for combining the loads also exist [101, 102].

Due to the very construction of the chirp signals, their frequency increases

linearly with time. This may be exploited to further regularize the inversion process,

i.e., we start the inversion process by considering only a portion of the total chirp

duration, arrive at an inverted profile, use the inverted profile as an initial guess to

the next round of inversion, where we increase the duration of the same chirp signal,

thus, gradually, bringing additional frequencies to bear on the inversion process (a,

so-called, continuation scheme). For example, for the site under study, we start with

the first 2 s of the signals, and progressively move up to 7 s duration, in increments

of 1 s. A similar idea was discussed in [58, 103].

160

After 1900 iterations, the misfit between the measured response and the com-

puted response becomes small enough, with no discernible update in the material

properties. The corresponding compressional and shear wave velocity profiles for the

Hornsby Bend site are shown in Fig. 4.19 (a constant mass density of ρ = 2000 kg/m3

is considered for the soil medium throughout the analysis).

x1 (m)

x3

(m)

−100 −90 −80 −70 −60 −50 −40 −30 −20 −10 0 10 20 30 40 50 60 70 80 90 100

−50

−40

−30

−20

−10

0

200

300

400

500

600

700

800

900

(a) cp (m/s)

x1 (m)

x3

(m)

−100 −90 −80 −70 −60 −50 −40 −30 −20 −10 0 10 20 30 40 50 60 70 80 90 100

−50

−40

−30

−20

−10

0

100

200

300

400

500

600

(b) cs (m/s)

Figure 4.19: Inverted profiles for cp and cs at iteration 1900.

4.4.2 Comparison with SASW

Next, we compare our shear wave velocity (cs) profile with that obtained via

the SASW method. The Spectral-Analysis-of-Surface-Waves method relies on the

dispersive nature of the Rayleigh wave velocity in layered media, i.e., the propaga-

tion speed of the surface waves depends on the frequency of the load [59]. Measuring

161

this wave speed for different frequencies in a field experiment, results in the exper-

imental dispersion curve. Next, a theoretical dispersion curve can be computed for

a homogeneous elastic layered medium. The material properties for each layer are

varied until a match is attained between the experimental and the theoretical disper-

sion curve (a comprehensive description of the SASW method can be found in [104]).

The method assumes that the dominant portion of the wave energy is transported

through Rayleigh waves, and disregards other wave types such as compressional and

shear waves. The SASW is capable of rendering only horizontally layered profiles

and only of the shear wave velocity. Despite its limitations it is widely used.

We performed three SASW experiments at our site: at the centerpoint (x1, x2) =

(0, 0) and close to the two end points of the domain (x1, x2) = (±90, 0). The cs

profile corresponding to the center point is shown in Fig. 4.20, whereas Fig. 4.21

compares the SASW profile with those obtained from the inversion process at the

x1 = −90 m, 0 m, and +90 m cross-sectional lines of the domain. In general, there

is good agreement between the two methods. Discrepancies may be attributed to

the three-dimensional nature of the physical problem. While in our study, we use a

two-dimensional model for the full-waveform inversion, the model is one-dimensional

for the SASW method. Whereas there may exist lateral property variability in the

actual physical problem, these effects are completely neglected in the SASW method,

and are only partially accounted for in the two-dimensional inversion process. We

also observe that while the SASW method predicts sharp profile changes in depth,

the inversion process yields profiles that vary gradually. Indeed, this is due to the

Tikhonov regularization scheme, which precludes high gradients in the material pro-

162

file while allowing smooth spatial variations.

We also compute time-history results corresponding to numerical simulation

of the medium based on profiles obtained from the inversion process and the SASW

method, and compare them with the actual field measurements at a few locations.

To this end, since the SASW method only yields the shear wave velocity of the soil

medium, we supplemented it with an estimation of the compressional wave velocity,

by assuming that the Poisson’s ratio decreases from 0.35 on the ground surface to

0.25 at a depth of −50 m. This allowed us to compute the response at the ground

surface using the SASW-rendered profile. Due to the small variability between the

three SASW profiles, we use their average in the time-history analysis for simplicity.

Displacement time-histories at x1 = −80 m, −40 m, +5 m, +35 m, and +70 m are

shown in Fig. 4.22. Excellent agreement can be observed between the time histories

computed based on the full-waveform inverted profile and the true recorded motion,

whereas the SASW-based time histories exhibit significant amplitude discrepancies.

We should mention, however, that since the full-waveform inversion process forces

time-history matching, this may not be a fair comparison.

x1 (m)

x3

(m)

−100 −90 −80 −70 −60 −50 −40 −30 −20 −10 0 10 20 30 40 50 60 70 80 90 100

−50

−40

−30

−20

−10

0

100

200

300

400

500

600

Figure 4.20: Inverted profile for cs via the SASW method.

163

0 100 200 300 400 500 600−50

−40

−30

−20

−10

0

cs (m/s)

x3

(m)

SASWInversion

(a) x1 = −90 m

0 100 200 300 400 500 600−50

−40

−30

−20

−10

0

cs (m/s)

x3

(m)

SASWInversion

(b) x1 = ±0 m

0 100 200 300 400 500 600−50

−40

−30

−20

−10

0

cs (m/s)

x3

(m)

SASWInversion

(c) x1 = +90 m

Figure 4.21: Shear wave velocity profiles obtained via SASW and full-waveform-basedinversion.

4.4.3 Comparison with cone penetration test (CPT) results

The cone-penetration-test (CPT) is an intrusive field experiment, which pro-

vides information on soil properties and the soil’s stratification. A rod with a cone-

shaped ending is forced into the ground at a constant rate, while two load cells

measure the required load that drives the rod into the ground. The first load cell

measures the force that acts directly on the cone and yields the cone resistance, which

is obtained by dividing the force over the cone’s area. The second load cell measures

the force that acts on the lateral sides of the rod, immediately past the cone, and

provides sleeve friction [105]. Unfortunately, CPT results cannot be correlated with

elastic properties of the soil. However, the cone resistance is an indicator of soil

stiffness, and the depth at which the cone cannot be forced further into the ground,

corresponds to a stiff layer. We present CPT results at the same site, and consider

mainly the qualitative information that the test provides.

164

−1 0 1 2 3 4 5 6 7 8−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredinversionSASW

(a) x1 = −80 m

−1 0 1 2 3 4 5 6 7 8−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredinversionSASW

(b) x1 = −40 m

−1 0 1 2 3 4 5 6 7 8−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredinversionSASW

(c) x1 = +5 m

−1 0 1 2 3 4 5 6 7 8−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredinversionSASW

(d) x1 = +35 m

−1 0 1 2 3 4 5 6 7 8−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredinversionSASW

(e) x1 = +70 m

Figure 4.22: Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion.

165

We performed CPT tests at four sample locations to investigate if they hit the

stiff zones predicted by the inverted profile. These locations are at the x1 = −80 m,

−50 m, −8 m, and +80 m cross-sectional lines of the domain. The cone resistance

along depth at these locations is shown in Fig. 4.23, along with the inverted cp and

cs profiles. It can be observed that the cone cannot be pushed any further once

it reaches a zone where the shear wave velocity is approximately around 400 m/s.

There is general agreement between the CPT results and the inverted profile. For

instance, at x1 = −80 m, the cone resistance and the shear wave velocity have the

same pattern in depth. At x1 = −50 m, cone resistance has a spike at a shallow

depth, reaches its minimum value at a depth of 5 m, and increases again after that.

We observe a similar trend for both the shear and the compressional wave velocities.

It is difficult to find a correlation at x1 = −8 m and +80 m.

4.5 Three-dimensional site characterization

In this section, we report the design and data processing of a field experi-

ment that we performed at the NEES@UCSB site in Garner Valley, CA. We use

the field data in our three-dimensional full-waveform inversion code, to obtain the

compressional wave velocity (cp) and shear wave velocity (cs) profiles of the site. The

full-waveform-inversion-based cs profile will then be compared against the profiling

obtained from the SASW method.

We use the T-Rex seismic vibrator of the NEES@UTexas site for applying

loads on the ground surface, and record the resulting ground motion by using 1 Hz

geophones. We use a chirp signal with characteristic parameters f0 = 3 Hz, and

166

0 200 400 600 800−15

−10

−5

0

wave velocity (m/s)

cscp

0 5 10 15 20

−15

−10

−5

0

cone resistance (MPa)

dep

th(m

)

(a) x1 = −80 m

0 200 400 600 800−15

−10

−5

0

wave velocity (m/s)

cscp

0 5 10 15 20

−15

−10

−5

0

cone resistance (MPa)

dep

th(m

)

(b) x1 = −50 m

0 200 400 600 800−15

−10

−5

0

wave velocity (m/s)

cscp

0 5 10 15 20

−15

−10

−5

0

cone resistance (MPa)

dep

th(m

)

(c) x1 = −8 m

0 200 400 600 800−15

−10

−5

0

wave velocity (m/s)

cscp

0 5 10 15 20

−15

−10

−5

0

cone resistance (MPa)

dep

th(m

)

(d) x1 = +80 m

Figure 4.23: Juxtaposition of CPT results and the inverted profiles.

167

k = 2.8 (see (4.28)), and a total active duration of 2.5 s. This parameterization

results in a loading with dominant frequencies between 3-10 Hz.

4.5.1 The experiment site and layout

The NEES Garner Valley Downhole Array research site in Southern California

is located in a narrow valley, within the Peninsular Ranges batholith, 23 km east of

Hemet, 20 km southwest of Palm Springs, and is just 7 km from the San Jacinto

fault, and 35 km from the San Andreas fault.

We consider a portion of the site of length and width 126 m × 68 m, and

40 m depth. The experiment layout is displayed in Fig. 4.24. The main grid lines

are 10 m apart. The sensors are shown with bullets and sources are indicated by

squares. Overall, we “shake” at 24 locations, and record the site’s response at 63

receiver locations. The experiment was performed on March 13th 2012 (Fig. 4.25).

Figure 4.24: The field experiment layout.

168

(a) (b)

(c) (d)

Figure 4.25: Garner Valley field experiment: (a) T-rex at the site; (b) a buriedgeophone; (c) instrumentation van; and (d) the site.

169

4.5.2 Pre-processing the field data

We use a sampling frequency of 200 Hz for digital data collection, which,

according to the Nyquist sampling theorem [100], is adequate, and prevents aliasing,

if the recorded data are contaminated by noise up to 100 Hz. Moreover, to reduce

the effects of ambient noise, we repeat each loading five times, and use the averaged

recording for inversion. We then process the signal to eliminate the parts of the

recorded motion with low signal to noise ratio, according to the procedure described

in Section 4.3.1.

Next, we present a subset of the recorded motion at select sensor locations.

Specifically, we consider the displacement time-history at four sensors, placed at

(x, y) = (30, 20), (30, 50), (70, 20), (70, 50) m, due to loading at (x, y) = (50, 35) m.

These sensors are shown with blue bullets in Fig. 4.26, and are labeled as “array

1”. The considered sensors are equidistant from the source. We also study the

recorded motion at sensors placed at (x, y) = (10, 20), (10, 50), (90, 20), (90, 50) m,

and label them as “array 2”, which are shown with green bullets in Fig. 4.26. The

displacement time-history at array 1 and array 2 sensors are shown in Figs. 4.27 and

4.28, respectively. Although there is general agreement between the time-history

plots for the equidistant sensors of each array, there are differences: indicative of the

heterogeneous character of the site.

4.5.3 Full-waveform inversion using field data

Based on the inversion framework discussed in the preceding chapter, we use

the collected field data to compute the compressional wave velocity cp and shear

170

Figure 4.26: Location of array 1 and array 2 sensors.

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(a) (x, y) = (30, 50) m

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(b) (x, y) = (70, 50) m

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(c) (x, y) = (30, 20) m

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(d) (x, y) = (70, 20) m

Figure 4.27: Time-history of vertical displacement at sensors in array 1.

171

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(a) (x, y) = (10, 50) m

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(b) (x, y) = (90, 50) m

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(c) (x, y) = (10, 20) m

0 1 2 3 4−2

−1

0

1

2x 10

−5

t (sec)

u(m

)

(d) (x, y) = (90, 20) m

Figure 4.28: Time-history of vertical displacement at sensors in array 2.

172

wave velocity cs profiles of the probed site. We consider a cubic (regular) domain

of length, width, and depth 126 × 68 × 40 m. A 10 m-thick PML is placed at the

truncation boundaries. For the PML parameters, we choose αo = 5, βo = 500 s 1,

and a quadratic profile for the attenuation functions, i.e., m = 2. The mass density

is considered to be ρ = 1760 kg/m3 for −2 m ≤ z ≤ 0, ρ = 1880 kg/m3 for

−4 m ≤ z ≤ −2 m, and ρ = 2000 kg/m3 for −40 m ≤ z ≤ −4 m, according to

prior investigations. The material properties at the interfaces ΓI are extended into

the PML. The interior and PML domains are discretized by quadratic hexahedral

spectral elements of size 2 m, and ∆t = 10−3 s. This leads to 3, 885, 648 state

unknowns, and 718, 566 material parameters.

Owing to linearity, we apply all the 24 loads simultaneously, and consider

their superimposed responses at each sensor location. We start the inversion process

with only a portion of the signal, and as the inversion process evolves, we increase the

temporal duration of the signal, thus, progressively introducing higher frequencies,

which allows for profile refinement. We use a smoothed version of the profile obtained

via the SASW method as the starting point of the inversion process.

After 4550 iterations, the misfit between the measured response and the com-

puted response becomes reasonably small, with no strong update in the material

profiles. The corresponding three-dimensional cp and cs profiles for the Garner Val-

ley site are shown in Fig. 4.29, whereas Figs. 4.30-4.32 show the cross-sectional

profiles of the velocity profiles at x = 10, 60, and 90 m. The profiles indicate that

the site under investigation is heterogeneous, revealing the presence of a soft layer

at a depth of about 25 m.

173

200 400 600

100 700

(a) cp (m/s)

200 300

100 400

(b) cs (m/s)

Figure 4.29: Inverted profiles for cp and cs at iteration 4550 (part of the domain isnot shown to aid material visualization in depth).

174

200 400 600

100 700

(a) cp (m/s)

200 300

100 400

(b) cs (m/s)

Figure 4.30: Inverted profiles for cp and cs at iteration 4550 (x = 10 m).

175

200 400 600

100 700

(a) cp (m/s)

200 300

100 400

(b) cs (m/s)

Figure 4.31: Inverted profiles for cp and cs at iteration 4550 (x = 60 m).

176

200 400 600

100 700

(a) cp (m/s)

200 300

100 400

(b) cs (m/s)

Figure 4.32: Inverted profiles for cp and cs at iteration 4550 (x = 90 m).

177

4.5.4 Profiling obtained from SASW

We performed field tests using the SASW method to obtain the shear wave

velocity (cs) profile of the site along three lines. The profiles are obtained along

x = 10, 60, and 100 m, which are marked with red lines in Fig. 4.33, and are

shown in Fig. 4.34. The cs profiles obtained from the full-waveform inversion (FWI)

at three different locations along these lines are also shown in the plots. There are

considerable differences between the SASW profile and the FWI profiles along the

x = 10 m and x = 60 m lines. However, there is general agreement between the

profiles along the x = 100 m line.

(a) (b)

Figure 4.33: Garner Valley experiment: (a) layout; and (b) SASW method testlocations.

Next, we compute the time-history response corresponding to a forward wave

simulation of the site, based on profiles obtained from full-waveform inversion and

the SASW method7 (Fig. 4.35), and compare them against the measured field re-

7We consider ν=0.3 to compute a cp profile for the SASW method in order to exercise theforward code. Furthermore, since the SASW profiles along the three lines are almost identical, weuse their average for time-history simulations.

178

0 50 100 150 200 250 300 350−40

−35

−30

−25

−20

−15

−10

−5

0

z(m

)

cs (m/s)

SASWFWI y = 0 mFWI y = 30 mFWI y = 60 m

(a) x = 10 m

0 50 100 150 200 250 300 350−40

−35

−30

−25

−20

−15

−10

−5

0

z(m

)

cs (m/s)

SASWFWI y = 0 mFWI y = 30 mFWI y = 60 m

(b) x = 60 m

0 50 100 150 200 250 300 350−40

−35

−30

−25

−20

−15

−10

−5

0

z(m

)

cs (m/s)

SASWFWI y = 0 mFWI y = 30 mFWI y = 60 m

(c) x = 100 m

Figure 4.34: Shear wave velocity profiles of the NEES site obtained via SASW andfull-waveform-based inversion (FWI).

179

sponse at select sensor locations. The sensors are placed along the x = 10, 60, 90, 100

m lines (Figs. 4.36-4.39). There is good agreement between the measured field

response and the response computed based on using the full-waveform inversion pro-

files. The agreement is significantly better for sensors located along the x = 90,

100 m lines (Figs. 4.38-4.39). We remark that the full-waveform-inversion-based cs

profile also agrees better with that of the SASW profile for the x = 100 m line (Fig.

4.34(c)). The time-histories at the sensors, computed based on the SASW profiles,

differ significantly from the recorded signals.

200 300

100 400

Figure 4.35: Inverted profile for cs via the SASW method.

180

0 0.5 1 1.5 2 2.5 3−2.5

−1.25

0

1.25

2.5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(a) y = +60 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(b) y = +50 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(c) y = +40 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(d) y = +30 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(e) y = +20 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(f) y = +10 m

Figure 4.36: Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion (x = +10m).

181

0 0.5 1 1.5 2 2.5 3−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(a) y = +60 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(b) y = +50 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(c) y = +40 m

0 0.5 1 1.5 2 2.5 3−2.5

−1.25

0

1.25

2.5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(d) y = +30 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(e) y = +20 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(f) y = +10 m

Figure 4.37: Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion (x = +60m).

182

0 0.5 1 1.5 2 2.5 3−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(a) y = +60 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(b) y = +50 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(c) y = +40 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(d) y = +30 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(e) y = +20 m

0 0.5 1 1.5 2 2.5 3−4

−2

0

2

4x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(f) y = +10 m

Figure 4.38: Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion (x = +90m).

183

0 0.5 1 1.5 2 2.5 3−2

−1

0

1

2x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(a) y = +60 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(b) y = +50 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(c) y = +40 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(d) y = +30 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(e) y = +20 m

0 0.5 1 1.5 2 2.5 3−5

−2.5

0

2.5

5x 10

−4

t (sec)

uz

(m)

measuredFWISASW

(f) y = +10 m

Figure 4.39: Comparison of measured surface displacement time-histories againstthose resulting from the SASW and full-waveform-based inversion (x = +100m).

184

4.6 Summary

We discussed recent advances in the development of a general and robust

methodology for geotechnical site characterization, based on full-waveform inversion,

using actual field data.

We considered two field experiments consistent with our code development:

firstly, we characterized the Hornsby Bend site in Austin, TX, by creating plane

strain field conditions and by assuming homogeneity lateral to a two-dimensional

slice. We discussed the design and post-processing needs of such field experiments,

so that collected records can be seamlessly integrated in the software toolchain.

Secondly, we reported the design of a field experiment aimed at the three-dimensional

characterization of the NEES@UCSB site in Garner Valley, CA, using the developed

three-dimensional inversion codes discussed in the preceding chapter.

Our results demonstrate the clear advantage of allowing for the imaging of

arbitrarily heterogeneous sites. Overall, our full-waveform-inversion-based site char-

acterization methodology seems robust and promising.

185

Chapter 5

Conclusions

5.1 Summary

We presented a robust methodology for site characterization using full-waveform

inversion. We complemented the theory and the numerical implementation with re-

sults derived from synthetic, as well as field data.

We developed PMLs for domain truncation of three-dimensional, arbitrarily

heterogeneous, elastic formations. The formulation is implemented in parallel, and

uses scalable algorithms, which makes it suitable for tackling large-scale problems.

Specifically, we use a displacement-stress formulation for the PML, coupled with a

standard displacement-only formulation for the PML. This hybrid treatment leads

to a cost-efficient computational scheme. We discuss several time-marching schemes,

which can be used a la carte, depending on the application: a) an extended Newmark

scheme for third-order in time, either unsymmetic or fully symmetric semi-discrete

forms; b) a standard implicit Newmark for second-order, unsymmetric semi-discrete

forms; and c) an explicit Runge-Kutta scheme for a first-order in time unsymmetric

system. We also discuss how our formulation can accommodate M-PML with simple

modifications. Using numerical experiments, we demonstrate stability and efficacy

of the proposed formulation.

186

Armed with a parallel state-of-the-art forward-wave-solver, we considered the

elastic inverse medium problem in three space dimensions. We cast the problem of

finding the distributed Lame parameters in an arbitrarily heterogeneous formation,

as a PDE-constraint optimization problem: elastic waves are used as probing agents

to interrogate the soil medium, and the response of the medium to these waves are

collected at receivers located on the ground surface. The inversion process relies

on minimizing a misfit between the collected response at receiver locations, and

a computed response based on a trial distribution of the Lame parameters. We

used strategies to alleviate ill-posedness, and lend algorithmic robustness to the

proposed inversion scheme. Specifically, we used (a) a regularization factor selection

and continuation scheme, where the regularization factor is adaptively computed

at each inversion iteration, based on a simple procedure; (b) a source-frequency

continuation scheme such that the inversion process evolves by using low-frequency

sources, and, gradually, we use sources with higher frequencies; and (c) a biasing

scheme for the λ search direction, such that, at early iterations of inversion, the

search direction for λ is biased based on that of µ. The latter strategy, in particular,

improves the reconstruction of the material profiles when simultaneous inversion of

the two Lame parameters is exercised. After verifying the accuracy of the computed

discrete gradients, by comparing them with directional finite differences, we reported

results demonstrating successful reconstruction of the Lame parameters for smooth

and sharp profiles, using both noise-free and highly-noisy data.

We also considered the inverse medium problem in two space dimensions,

followed by presenting a practical procedure to accommodate three-dimensional field

187

data into two-dimensional codes. We then presented inversion results pertaining

to a field experiment at the Hornsby Bend in Austin, TX, and compared them

against profiles obtained from the non-invasive SASW, and from invasive CPT tests.

Lastly, we used the methodology described in Chapter 3 for the three-dimensional

site characterization of the NEES@UCSB site in Garner Valley, CA. Overall, the

framework discussed in this study seems robust, practical, and promising.

5.2 Future directions

We suggest future directions that were outside the scope of this disserta-

tion, but are perceived as significant steps toward improving the proposed inversion

methodology.

• Validation: in an attempt to validate the proposed methodology, it is important

for the full-waveform-inversion-based profiles, obtained from the field experi-

ments, to be compared against invasive techniques that can also provide wave

velocity profiles, such as borehole methods.

• Uncertainty quantification: it is desirable to quantify how much confidence one

has in the inverted profiles. The Bayesian framework discussed in [106] seems

to be applicable to the methodology discussed in this work for quantifying

uncertainty in the inverted material profiles.

• Enhanced physics for wave simulation: reliable simulation of wave propagation

in soils is remarkably challenging. Even though soil is a lossy, porous medium

that is, oftentimes, partially saturated, it is usually idealized as an elastic solid.

188

Using models that can capture the complex physics of the soil more realistically,

is a significant improvement toward high-fidelity subsurface imaging.

• Multi-scale approach for inversion: in order to alleviate the ill-posedness in-

herent in inverse problems, and to minimize the risk of local minima trapping,

using a multi-scale approach, whereby the inversion process in carried out on

a sequence of finer grids, seems to be a viable approach [47].

189

Appendices

190

Appendix A

Submatrix definitions

Subscripts in the shape functions indicate derivatives.

A.1 Submatrices in equation (2.31)

KRD =

ΩRD

Kxx Kxy Kxz

Kyx Kyy Kyz

Kzx Kzy Kzz

dΩ,

Kxx = (λ+ 2µ)ΦxΦTx + µ(ΦyΦ

Ty +ΦzΦ

Tz ),

Kxy = λΦxΦTy + µΦyΦ

Tx ,

Kxz = λΦxΦTz + µΦzΦ

Tx ,

Kyx = λΦyΦTx + µΦxΦ

Ty ,

Kyy = (λ+ 2µ)ΦyΦTy + µ(ΦxΦ

Tx +ΦzΦ

Tz ),

Kyz = λΦyΦTz + µΦzΦ

Ty ,

Kzx = λΦzΦTx + µΦxΦ

Tz ,

Kzy = λΦzΦTy + µΦyΦ

Tz ,

Kzz = (λ+ 2µ)ΦzΦTz + µ(ΦxΦ

Tx +ΦyΦ

Ty ). (A.1a)

MRD =

ΩRD

ρ diag(ΦΦT ,ΦΦT ,ΦΦT ) dΩ. (A.2a)

Mi =

ΩRD

i ρ diag(ΦΦT ,ΦΦT ,ΦΦT ) dΩ, i = a, b, c, d. (A.2b)

Ni =

ΩPML

i diag(ΨΨT ,ΨΨT ,ΨΨT , 2ΨΨT , 2ΨΨT , 2ΨΨT ) dΩ, i = a, b, c, d. (A.2c)

191

Aiu =

ΩPML

ΦxΨ

T λiyz ΦyΨ

T λixz ΦzΨ

T λixy

ΦyΨT λi

xz ΦxΨT λi

yz ΦzΨT λi

xy

ΦzΨT λi

xy ΦxΨT λi

yz ΦyΨT λi

xz

dΩ,

i = e, w, p, λejk = αjαk, λp

jk = αjβk + βjαk, λwjk = βjβk, j, k = x, y, z.

(A.3a)

Ail =

ΩPML

Ax1 Ax2 Ax3 Ax4 Ax5

Ay1 Ay2 Ay3 Ay4 Ay6

Az1 Az2 Az3 Az5 Az6

dΩ, (A.4a)

Ax1 = (λ+ 2µ)ΦxΨT λi

yz, Ay1 = λΦyΨT λi

xz, Az1 = λΦzΨT λi

xy,

Ax2 = λΦxΨT λi

yz , Ay2 = (λ+ 2µ)ΦyΨT λi

xz, Az2 = λΦzΨT λi

xy,

Ax3 = λΦxΨT λi

yz , Ay3 = λΦyΨT λi

xz, Az3 = (λ+ 2µ)ΦzΨT λi

xy,

Ax4 = 2µΦyΨT λi

xz, Ay4 = 2µΦxΨT λi

yz , Az5 = 2µΦxΨT λi

yz ,

Ax5 = 2µΦzΨT λi

xy, Ay6 = 2µΦzΨT λi

xy, Az6 = 2µΦyΨT λi

xz ,

λejk = αjαk, λp

jk = αjβk + βjαk, λwjk = βjβk,

i = e, w, p, j, k = x, y, z. (A.5a)

fRD =

ΓRD

N

Φ gx(x, t)Φ gy(x, t)Φ gz(x, t)

dΓ +

ΩRD

Φ bx(x, t)Φ by(x, t)Φ bz(x, t)

dΩ. (A.6a)

A.2 Submatrices for the symmetric PML formulation

Ni =

ΩPML

i

λ+µµ(3λ+2µ)ΨΨT −λ

2µ(3λ+2µ)ΨΨT −λ2µ(3λ+2µ)ΨΨT

−λ2µ(3λ+2µ)ΨΨT λ+µ

µ(3λ+2µ)ΨΨT −λ2µ(3λ+2µ)ΨΨT

−λ2µ(3λ+2µ)ΨΨT −λ

2µ(3λ+2µ)ΨΨT λ+µµ(3λ+2µ)ΨΨT

1µΨΨT

1µΨΨT

1µΨΨT

dΩ,

i = a, b, c, d.(A.7a)

192

Ai =

ΩPML

ΦxΨ

T λiyz ΦyΨ

T λixz ΦzΨ

T λixy

ΦyΨT λi

xz ΦxΨT λi

yz ΦzΨT λi

xy

ΦzΨT λi

xy ΦxΨT λi

yz ΦyΨT λi

xz

dΩ,

i = e, w, p,

λejk = αjαk, λp

jk = αjβk + βjαk, λwjk = βjβk, j, k = x, y, z. (A.8a)

A.3 Submatrices for M-PML

Aiu =

ΩPML

ΦxΨ

T λiyz ΦyΨ

T λixz ΦzΨ

T λixy

ΦyΨT λi

xz ΦxΨT λi

yz ΦzΨT λi

xy

ΦzΨT λi

xy ΦxΨT λi

yz ΦyΨT λi

xz

+

ΩPML

ΦΨT ∂λiyz

∂xΦΨT ∂λi

xz

∂yΦΨT ∂λi

xy

∂z

ΦΨT ∂λixz

∂yΦΨT ∂λi

yz

∂xΦΨT ∂λi

xy

∂z

ΦΨT ∂λixy

∂zΦΨT ∂λi

yz

∂xΦΨT ∂λi

xz

∂y

dΩ,

i = e, w, p, λejk = αjαk, λp

jk = αjβk + βjαk, λwjk = βjβk, j, k = x, y, z.

(A.9a)

A.4 Discretization of the control problems

In Section 3.2.1.3, we discussed the λ- and µ-control problems. In this part,we consider their spatial discretization. We use the basis function Φ for the spatialdiscretization of w(x, t) and u(x, t), and χ is the basis function for discretizing λ(x)and µ(x). For instance, if we approximate λ(x) with λh(x), then λh(x) = χTλ, whereλ comprises the vector of nodal values for λ. In the following, subscripts in the shapefunctions indicate derivatives, and uh = (uT

x ,uTy ,u

Tz )

T and wh = (wTx ,w

Ty ,w

Tz )

T isthe vector of discrete values of the state and adjoint variables, respectively. Accord-ingly

M =

ΩRD

χχT dΩ. (A.10a)

193

For Tikhonov regularization:

gλreg =

ΩRD

(χxχTx + χyχ

Ty + χzχ

Tz )λ dΩ, (A.10b)

gµreg =

ΩRD

(χxχTx + χyχ

Ty + χzχ

Tz )µ dΩ. (A.10c)

For Total Variation regularization:

gλreg =

ΩRD

(χxχTx + χyχ

Ty + χzχ

Tz )λ

(λT (χxχ

Tx + χyχ

Ty + χzχ

Tz )λ+ ǫ

) 1

2

dΩ, (A.10d)

gµreg =

ΩRD

(χxχTx + χyχ

Ty + χzχ

Tz )µ

(µT (χxχ

Tx + χyχ

Ty + χzχ

Tz )µ+ ǫ

) 1

2

dΩ. (A.10e)

Moreover,

gλmis = −

∫ T

0

ΩRD

χ (ΦTxwx +ΦT

ywy +ΦTz wz)(Φ

Txux +ΦT

y uy +ΦTz uz) dΩ dt,

(A.10f)

gµmis = −

∫ T

0

ΩRD

χ(2 (ΦT

xwx ΦTxux +ΦT

ywy ΦTy uy +ΦT

z wz ΦTz uz)

+ (ΦTywx +ΦT

xwy)(ΦTxuy +ΦT

y ux) + (ΦTz wx +ΦT

xwz)(ΦTxuz +ΦT

z ux)

+ (ΦTz wy +ΦT

ywz)(ΦTy uz +ΦT

z uy))dΩ dt. (A.10g)

In (A.10a), upon using spectral elements with LGL quadrature rule, M becomesdiagonal; thus, its inverse can be computed easily.

A.5 Submatrices in equation (4.3)

The two-dimensional PML formulation discussed in Chapter 4, results in thesemi-discrete form (4.3), with the following definition for the system matrices:

194

M2D =

[MRD + Ma 0

0 −Na

], C2D =

[Mb Ae

ATe −Nb

], (A.11a)

K2D =

[KRD + Mc Ap

ATp −Nc

], (A.11b)

where a bar denotes matrix extension to encompass all the displacement degrees-of-freedom. The submatrices in (A.11) are defined below:

KRD =

ΩRD

[(λ+ 2µ)ΦxΦ

Tx + µΦyΦ

Ty λΦxΦ

Ty + µΦyΦ

Tx

λΦyΦTx + µΦxΦ

Ty (λ+ 2µ)ΦyΦ

Ty + µΦxΦ

Tx

]dΩ, (A.12a)

MRD =

ΩRD

ρ diag(ΦΦT ,ΦΦT ) dΩ, (A.12b)

Mi =

ΩRD

i ρ diag(ΦΦT ,ΦΦT ) dΩ, i = a, b, c, (A.12c)

Ni =

ΩPML

i

λ+2µ4µ(λ+µ)ΨΨT −λ

4µ(λ+µ)ΨΨT

−λ4µ(λ+µ)ΨΨT λ+2µ

4µ(λ+µ)ΨΨT

1µΨΨT

dΩ, i = a, b, c, (A.12d)

Ai =

ΩPML

[ΦxΨ

T λiy ΦyΨ

T λix

ΦyΨT λi

x ΦxΨT λi

y

]dΩ, i = e, p, λe

j = αj , λpj = βj , j = x, y,

(A.12e)

where a = αxαy, b = αxβy + αyβx, and c = βxβy.

195

Appendix B

Time-integration schemes

B.1 Fourth-order Runge-Kutta method

In Section 2.3.2 we discussed various time marching schemes for integrating (2.32). Our

preferred scheme is the explicit 4th-order Runge-Kutta method (RK-4), which is outlined below.

Upon using spectral elements, with Legendre-Gauss-Lobatto quadrature rule, the mass-like

matrix M becomes diagonal; therefore, its inverse can be readily computed. We define the following

variables:

C = M−1 C, K = M−1 K, (B.1a)

G = M−1 G, f = M−1 f . (B.1b)

Using the above notation, (2.36) becomes:

d

dt

x0

x1

x2

=

0 I 00 0 I

−G −K −C

x0

x1

x2

+

00

f

. (B.2)

The explicit RK-4 scheme entails computing the following vectors:

196

k10 = xn1 ,

k11 = xn2 ,

k12 = −Cxn2 − Kxn

1 − Gxn0 + fn,

k20 = xn1 +

∆t

2k11,

k21 = xn2 +

∆t

2k12,

k22 = −C(xn2 +

∆t

2k12)− K(xn

1 +∆t

2k11)− G(xn

0 +∆t

2k10) + fn+

1

2 ,

k30 = xn1 +

∆t

2k21,

k31 = xn2 +

∆t

2k22,

k32 = −C(xn2 +

∆t

2k22)− K(xn

1 +∆t

2k21)− G(xn

0 +∆t

2k20) + fn+

1

2 ,

k40 = xn1 +∆t k31,

k41 = xn2 +∆t k32,

k42 = −C(xn2 +∆t k32)− K(xn

1 +∆t k31)− G(xn0 +∆t k30) + fn+1.

Finally, the solution at time step (n+ 1) can be updated via

x0

x1

x2

n+1

=

x0

x1

x2

n

+∆t

6

k10 + 2 k20 + 2 k30 + k40

k11 + 2 k21 + 2 k31 + k41

k12 + 2 k22 + 2 k32 + k42

. (B.4)

B.2 Extended Newmark method

In this part, we are concerned with the time integration of the following semi-discrete

equation, discussed in Sections 2.3.2 and 2.5:

197

Md+Cd+Kd+Gd = f , (B.5a)

d =

∫ t

0

d(τ)|PML dτ. (B.5b)

We discuss an extension of the Newmark method [107] for time integration of this equation.

The scheme is implicit, and can be applied to problems with either symmetric, or unsymmetric

matrices. We start with Taylor series-like expansion of the following quantities

dn+1 = dn +∆t dn +∆t2

2dn + (

1

6− α)∆t3 dn + α ∆t3 dn+1, (B.6a)

dn+1 = dn +∆t dn + (1

2− β)∆t2 dn + β ∆t2 dn+1, (B.6b)

dn+1 = dn + (1 − γ)∆t dn + γ ∆t dn+1, (B.6c)

where ∆t denotes the time step, superscripts (n) and (n+ 1) indicate current and next time steps,

β and γ are the classic Newmark parameters, and α is a new parameter. Substitution of (B.6) in

(B.5) at the (n+ 1)th

time step, results in the following linear system of equations

K dn+1 = Rn+1, (B.7a)

198

where

K = M+ γ ∆t C+ β ∆t2 K+ α ∆t3 G, (B.7b)

Rn+1 = fn+1

−C[dn + (1− γ) ∆t dn

]

−K

[dn +∆t dn + (

1

2− β) ∆t2 dn

]

−G

[dn +∆t dn +

∆t2

2dn + (

1

6− α) ∆t3 dn

]. (B.7c)

Upon solving for dn+1 from (B.7a), dn+1, dn+1, and dn+1 can be updated using (B.6). Average-,

and linear-acceleration schemes correspond to taking (α, β, γ) equal to ( 112 ,

14 ,

12 ), and ( 1

24 ,16 ,

12 ), re-

spectively. Numerical results reported in Section 2.7.3 were computed using the average-acceleration

scheme.

B.3 The adjoint problem time-integration scheme

We outline the explicit 4th-order Runge-Kutta method (RK-4) for the reverse time-integration

of the adjoint problem. Upon using spectral elements, with Legendre-Gauss-Lobatto (LGL) quadra-

ture rule, the mass-like matrix M becomes diagonal; therefore, its inverse can be readily computed.

We use the following notation:

C = C M−1, K = K M−1, (B.8a)

G = G M−1, f = M−1 fadj. (B.8b)

Using (B.8), (3.14) becomes:

199

d

dt

y0

y1

y2

=

0 I 00 0 I

GT −KT CT

y0

y1

y2

+

00

f

. (B.9)

The scheme entails computing the following vectors:

k10 = yn1 ,

k11 = yn2 ,

k12 = Cyn2 − Kyn

1 + Gyn0 + fn,

k20 = yn1 −

∆t

2k11,

k21 = yn2 −

∆t

2k12,

k22 = C(yn2 −

∆t

2k12)− K(yn

1 −∆t

2k11) + G(yn

0 −∆t

2k10) + fn−

1

2 ,

k30 = yn1 −

∆t

2k21,

k31 = yn2 −

∆t

2k22,

k32 = C(yn2 −

∆t

2k22)− K(yn

1 −∆t

2k21) + G(yn

0 −∆t

2k20) + fn−

1

2 ,

k40 = yn1 −∆t k31,

k41 = yn2 −∆t k32,

k42 = C(yn2 −∆t k32)− K(yn

1 −∆t k31) + G(yn0 −∆t k30) + fn−1.

Finally, the solution at time step (n− 1) can be updated via

y0

y1

y2

n−1

=

y0

y1

y2

n

− ∆t

6

k10 + 2 k20 + 2 k30 + k40

k11 + 2 k21 + 2 k31 + k41

k12 + 2 k22 + 2 k32 + k42

. (B.11)

200

Appendix C

Gradient of a functional

The gradient of a functional F : H → R, where H is a Hilbert space, is defined as the

Riesz-representation of the derivative F′(q)(q), such that

( G(q), q )H = F′(q)(q) ∀q ∈ H, (C.1)

where G denotes the gradient, and we use the following notation for the Gateaux derivative of F at

q in a direction q:

F′(q)(q) = lim

h→0

F(q+ hq)− F(q)

h. (C.2)

With this definition, it is not possible to talk about the gradient, without specifying the inner-

product utilized to represent the derivative [108].

201

Appendix D

On the third discrete optimality condition

We discuss the derivation of the discrete control equations, i.e., of the third discrete opti-

mality condition, given in (4.19) and (4.20). We take the derivative of L with respect to λ and µ

over the interior domain only, since the values of the Lame parameters at the interface nodes are

extended into the PML domain, without any variation along the direction of projection [57]. This

assumption greatly simplifies the derivation and implementation of the control equations.

The λ control problem

Equation (4.1a) governs the interior domain, denoted by ΩRD, where λ, µ only contribute

to terms in the K matrix of (4.3), which contributes to Q in (4.7). We denote the part of K that

belongs to the interior domain by KRD: it is the stiffness matrix of the interior problem, and is

given by [33]:

KRD =

ΩRD

[(λTχ+ 2 µTχ) Φx1

ΦTx1

+ (µTχ) Φx3ΦT

x3. . .

(λTχ) Φx3ΦT

x1+ (µTχ) Φx1

ΦTx3

. . .

. . . (λTχ) Φx1ΦT

x3+ (µTχ) Φx3

ΦTx1

. . . (λTχ+ 2 µTχ) Φx3

ΦTx3

+ (µTχ) Φx1ΦT

x1

]dΩ, (D.1)

where χ is the vector of interpolation functions for the Lame parameters, Φ are the displacement

interpolants, and subscripts x1, x3 denote differentiation with respect to x1, and x3, respectively.

Taking the derivative of KRD with respect to λ, yields

202

∂KRD

∂λ=

ΩRD

[Φx1

ΦTx1χ Φx1

ΦTx3χ

Φx3ΦT

x1χ Φx3

ΦTx3χ

]dΩRD, (D.2)

which are the building blocks of ∂

∂λ(pT Q d) in (4.19), thus enabling its computation.

The µ control problem

In a way similar to what we did above, we take the derivative of KRD with respect to µ to

obtain

∂KRD

∂µ=

ΩRD

[2 Φx1

ΦTx1

+Φx3ΦT

x3

]χ Φx3

ΦTx1χ

Φx1ΦT

x3χ

[2 Φx3

ΦTx3

+Φx1ΦT

x1

dΩRD, (D.3)

thus allowing the computation of ∂∂µ(pT Q d) in (4.20).

203

Appendix E

On the singular convolution integral (4.21)

Due to the arrival of Raleigh waves, the integrand in (4.21) exhibits singularity. In this

part, we consider the integral in the Cauchy principal value sense, investigate the type of singularity,

and comment on the numerical evaluation of the integral. We repeat (4.21) with a small change in

notation:

u2D(x, t) =

∫ t

0

f2D(t− t′) G2D(x, t′) dt′, (E.1)

where G2D(x, t) is the Lamb Green’s function:

G2D(x, t) =cs

πµ|x|

0, τ < a−(2τ2−1)2

√τ2−a2

(2τ2−1)4+16τ4(τ2−a2)(1−τ2) , a ≤ τ ≤ 1−√τ2−a2

(2τ2−1)2−4τ2√τ2−a2

√τ2−1

, τ > 1,

where cs is shear wave velocity, µ is shear modulus, τ = tcs|x| is dimensionless time, a =

√1−2ν2−2ν

indicates the ratio of shear wave velocity to that of the compressional wave, and ν is Poisson’s

ratio.

In this section, we consider τ > 1, which contains the arrival time of the Rayleigh wave,

denoted by tR. Evaluation of (E.1) when τ ≤ 1 is straightforward. We split the integral into two

parts, the regular part and the singular part

204

u2D(x, t) =

t\B(tR,ǫ)

f2D(t− t′) G2D(x, t′) dt′ +

B(tR,ǫ)

f2D(t− t′) G2D(x, t′) dt′, (E.2)

where the singularity is isolated inside the ball B(tR, ǫ) = (tR − ǫ, tR + ǫ), where ǫ > 0 is “small”.

We focus on the singular integral. Since ǫ is “small”, we use the following approximation

B(tR,ǫ)

f2D(t− t′) G2D(x, t′) dt′ ≈ f2D(t− tR)

B(tR,ǫ)

G2D(x, t′) dt′. (E.3)

Next, we evaluate the following integral in the Cauchy principal value sense

Is = limǫ→0

B(tR,ǫ)

G2D(x, t′) dt′, (E.4)

where

G2D(x, t′) =cs

πµ|x|−√τ2 − a2

(2τ2 − 1)2 − 4τ2√τ2 − a2

√τ2 − 1

. (E.5)

We rewrite the singular kernel (E.5) in the following form

G2D(x, t′) =Q

L−R=

Q(L+R)

(L−R)(L +R), (E.6)

where

Q =−csπµ|x|

√τ2 − a2, (E.7a)

L = (2τ2 − 1)2, (E.7b)

R = 4τ2√τ2 − a2

√τ2 − 1. (E.7c)

205

We investigate the case when L − R in (E.6) vanishes, i.e., (L − R)(L + R) = 0. We exploit the

following representation

(L−R)(L+R) = α(τ2 − ξ21)(τ2 − ξ22)(τ

2 − ξ23), (E.8)

where ξ2i , i = 1, 2, 3, are the roots of the corresponding polynomial on the left-hand-side, and α is

a scaling parameter. We associate ξ3 to the arrival time of the Rayleigh wave, and rewrite (E.6),

using (E.8), as follows

G2D(x, t′) =Q(L+R)

α(τ2 − ξ21)(τ2 − ξ22)(τ + ξ3)(τ − ξ3)

=φ(τ)

τ − ξ3, (E.9)

where

φ(τ) =Q(L+R)

α(τ2 − ξ21)(τ2 − ξ22)(τ + ξ3)

. (E.10)

We remark that the above representation is considered in B(tR, ǫ). Thus, φ(τ) is continuous, and

may be approximated with the following second order polynomial

φ(τ) = φ(t cs|x| ) ≈ a0 + a1t+ a2t

2, t ∈ B(tR, ǫ). (E.11)

Thus, (E.4) becomes

Is ≈|x|cs

limǫ→0

∫ tR+ǫ

tR−ǫ

a0 + a1t+ a2t2

t− tRdt. (E.12)

206

It can be shown that the above integral vanishes when evaluated in the Cauchy principal value

sense, i.e., the singularity is “weak”.

Next, we focus on the numerical evaluation of (E.1). In computer arithmetic, one cannot

approach infinitely close to the singular point. Thus, our strategy is to approximate the singular

integral analytically, inside a larger ball. To make the idea precise, let δ > ǫ, andB(tR, δ) ⊃ B(tR, ǫ),

where δ has the same order of magnitude that the time discretization has in the numerical evaluation

of the regular integral. Let t1 = inf B(tR, δ), and t2 = supB(tR, δ). Then, along the same lines, we

have

u2D(x, t) =

t\B(tR,δ)

f2D(t− t′) G2D(x, t′) dt′ +

B(tR,δ)

f2D(t− t′) G2D(x, t′) dt′. (E.13)

We define

Is =

B(tR,δ)

f2D(t− t′) G2D(x, t′) dt′. (E.14)

Thus,

Is =

∫ tR−ǫ

t1

· · ·+∫

B(tR,ǫ)

· · ·+∫ t2

tR+ǫ

· · · , (E.15)

where · · · represents the integrand in (E.14). The second integral vanishes when evaluated in the

Cauchy principal value sense, as was shown earlier. It can be shown that the remaining terms can

be approximated as

207

Is ≈ f2D(t− tR)|x|cs

ln| t2 − tR

t1 − tR|(b0 + b1 tR + b2 t2R)

+ (t2 − t1)

(b1 + b2 (tR +

t2 + t12

)

), (E.16)

where b0, b1, and b2 are the coefficients of the polynomial that approximates φ(τ) in B(tR, δ). When

dealing with floating point arithmetic, the best approximation occurs when tR = 12 (t1 + t2), where

the logarithmic term vanishes.

208

Appendix F

On the spatial integration of (4.25)

We assume that the line load is applied along the x2 axis with a total length of 2L, and

that the observer is positioned along the x1 axis. We make the following change of variable to ease

the analytical integration:

G2L(x1, t) =

∫ L

−L

G3D(x1, x2, t) dx2 = 2

∫ rL

x1

G3D(r, t)r√

r2 − x21

dr, (F.1)

where, r =√x21 + x2

2, rL =√x21 + L2, and the Green’s function G3D is given as:

G3D(r, t) = B(r)

0 τ < a

u1 a ≤ τ ≤ 1

u∗2 1 ≤ τ ≤ ξ3

1 τ > ξ3

. (F.2)

We have the following definitions:

B(r) =1

rB∗ =

1

r

1− ν

2πµ, τ = t

csr,

u1 =1

2− 1

2

3∑

i=3

Ai√|τ2 − ξ2i |

, u∗2 = 1− A3√

ξ23 − τ2,

a =

√1− 2ν

2− 2ν, Ai =

(1− 2ξ2i )2√|a2 − ξ2i |

4(ξ2i − ξ2j )(ξ2i − ξ2k)

,

209

where ν is Poisson’s ration, µ is shear modulus, cs is shear wave velocity, and ξi are the roots of

the Rayleigh function:

R(ξ2) = (2ξ2 − 1)2 + 4√ξ2 − a2

√ξ2 − 1 = 0.

The above relations are valid when ν < 0.2631. In such a case, all roots are real and satisfy

0 < ξ21 < ξ22 < a2 < 1 < ξ23 [98]. Making use of Heaviside functions, we rewrite (F.2) as follows

G3D(r, t) = B(r)

−u1H

(r − tcs

a

)+ (u1 − u∗

2)H

(r − tcs

1

)+

(u∗2 − 1)H

(r − tcs

ξ3

)+ 1

. (F.3)

Substituting (F.3) into (F.1) yields

G2L(x1, t) = 2 (I1 + I2 + I3 + I4), (F.4)

where

I1 =1

2B∗

∫ rL

x1

−H(r − tcsa)√

r2 − x21

+

3∑

i=1

∫ rL

x1

Ai r H(r − tcsa)

ξi

√(t2c2sξ2i

− r2)(r2 − x21)

dr,

I2 =1

2B∗

∫ rL

x1

−H(r − tcs1 )√

r2 − x21

−2∑

i=1

∫ rL

x1

Ai r H(r − tcs1 )

ξi

√(t2c2sξ2i

− r2)(r2 − x21)

dr

+

∫ rL

x1

A3 r H(r − tcs1 )

ξ3

√(t2c2sξ23

− r2)(r2 − x21)

dr,

I3 = B∗∫ rL

x1

−A3 r H(r − tcsξ3

)

ξ3

√(t2c2sξ23

− r2)(r2 − x21)

dr,

210

I4 = B∗∫ rL

x1

dr√r2 − x2

1

.

It is easy to verify that I1, I2, I3, and I4 consist of the following integrals with the corresponding

closed-form solution

F1 =

∫r H(r − tcs

λ)√

(r2 − t2c2sξ2

)(r2 − x21)

dr =1

2H

(r − tcs

λ

ln

(r2 − x21) + (r2 − t2c2s

ξ2) + 2

√(r2 − t2c2s

ξ2)(r2 − x2

1)

(t2c2sλ2 − x2

1) + (t2c2sλ2 − t2c2s

ξ2) + 2

√(t2c2sλ2 − t2c2s

ξ2)(

t2c2sλ2 − x2

1)

,

F2 =

∫r H(r − tcs

λ)√

−(r2 − t2c2sξ2

)(r2 − x21)

dr

= −1

2H

(r − tcs

λ

)tan−1

(t2c2sξ2− r2) + (x2

1 − r2)

2√(t2c2sξ2− r2)(r2 − x2

1)

+1

2H

(r − tcs

λ

)tan−1

(t2c2sξ2− t2c2s

λ2 ) + (x21 −

t2c2sλ2 )

2√(t2c2sξ2− t2c2s

λ2 )(t2c2sλ2 − x2

1)

,

F3 =

∫ √r2 − x2

1dr = ln(r +√r2 − x2

1),

F4 =

∫H

(r − tcs

λ

)√r2 − x2

1 dr = H

(r − tcs

λ

)

ln(r +√r2 − x2

1)

ln( tcsλ

+

√t2c2sλ2 − x2

1)

,

where λ takes the value of a, 1, or ξ3. Clearly, by exploiting the above closed-form expressions,

(4.25) can be evaluated accurately and efficiently.

211

Bibliography

[1] J. Bielak, R. W. Graves, K. B. Olsen, R. Taborda, L. Ramrez-Guzmn, S. M. Day, G. P. Ely,

D. Roten, T. H. Jordan, P. J. Maechling, J. Urbanic, Y. Cui, and G. Juve. The shakeout

earthquake scenario: Verification of three simulation sets. Geophysical Journal International,

180(1):375–404, 2010.

[2] U. Basu and A. K. Chopra. Perfectly matched layers for transient elastodynamics of un-

bounded domains. Int. J. Numer. Meth. Engng., 59:1039–1074, 2004.

[3] A. Fathi and V. Lotfi. Effects of reservoir length and rigorous fluid-foundation interaction on

the frequency response functions of concrete gravity dams. Dam Engineering, XXI(3):213–

233, January 2011.

[4] I. Epanomeritakis, V. Akcelik, O. Ghattas, and J. Bielak. A Newton-CG method for large-

scale three-dimensional elastic full-waveform seismic inversion. Inverse Problems, 24(3):034015,

2008.

[5] C. Jeong, L.F. Kallivokas, S. Kucukcoban, W. Deng, and A. Fathi. Maximization of wave

motion within a hydrocarbon reservoir for wave-based enhanced oil recovery. Journal of

Petroleum Science and Engineering, doi:10.1016/j.petrol.2015.03.009, 2015.

[6] P. M. Karve, S. Kucukcoban, and L. F. Kallivokas. On an inverse source problem for en-

hanced oil recovery by wave motion maximization in reservoirs. Computational Geosciences,

doi:10.1007/s10596-014-9462-7, 2014.

212

[7] P. M. Karve, , and L. F. Kallivokas. Wave energy focusing to subsurface poroelastic forma-

tions to promote oil mobilization. Geophysical Journal International, doi:10.1093/gji/ggv133,

2015.

[8] L.F. Kallivokas, A. Fathi, S. Kucukcoban, K.H. Stokoe II, J. Bielak, and O. Ghattas. Site

characterization using full waveform inversion. Soil Dynamics and Earthquake Engineering,

47:62 – 82, 2013.

[9] I. Harari, M. Slavutin, and E. Turkel. Analytical and numerical studies of a finite element

PML for the Helmholtz equation. Journal of Computational Acoustics, 8(1):121–137, 2000.

[10] D. Rabinovich, D. Givoli, and E. Becache. Comparison of high-order absorbing boundary

conditions and perfectly matched layers in the frequency domain. International Journal for

Numerical Methods in Biomedical Engineering, 26(10):1351–1369, 2010.

[11] F. Collino and P. B. Monk. Optimizing the perfectly matched layer. Computer Methods in

Applied Mechanics and Engineering, 164:157–171, 1998.

[12] P. G. Petropoulos. On the termination of the perfectly matched layer with local absorbing

boundary conditions. Journal of Computational Physics, 143(2):665–673, 1998.

[13] J.-P. Berenger. A perfectly matched layer for the absorption of electromagnetic waves.

Journal of Computational Physics, 114:185–200, 1994.

[14] W. C. Chew, J. M. Jin, and E. Michielssen. Complex coordinate system as a generalized

absorbing boundary condition. In Proc. 13th Annu. Rev. of Prog. Appl. Comp. Electromag.,

pages 909–914, Monterey,CA, 17-21 Mar. 1997. Vol. 2.

[15] W. C. Chew and W. H. Weedon. A 3D perfectly matched medium from modified Maxwell’s

equations with stretched coordinates. Micro. Opt. Tech. Lett., 7:599–604, 1994.

213

[16] F. L. Teixeira and W. C. Chew. Complex space approach to perfectly matched layers: a

review and some new developments. Int. J. Numer. Model., 13:441–455, 2000.

[17] W. C. Chew and Q. H. Liu. Perfectly matched layers for elastodynamics: a new absorbing

boundary condition. Journal of Computational Acoustics, 4(4):341–359, 1996.

[18] F. Collino and C. Tsogka. Application of the perfectly matched absorbing layer model to the

linear elastodynamic problem in anisotropic heterogeneous media. Geophysics, 66(1):294–

307, February 2001.

[19] F. Q. Hu. On absorbing boundary conditions for linearized Euler equations by a perfectly

matched layer. Journal of Computational Physics, 129:201–219, 1996.

[20] Y. Q. Zeng, J. Q. He, and Q. H. Liu. The application of the perfectly matched layer in

numerical modeling of wave propagation in poroelastic media. Geophysics, 66(4):1258–1266,

2001.

[21] S. D. Gedney. An anisotropic perfectly matched layer-absorbing medium for the truncation

of FDTD lattices. IEEE Transactions on Antennas and Propagation, 44(12):1630–1639,

December 1996.

[22] B. Gustafsson, H.-O. Kreiss, and J. Oliger. Time dependent problems and difference methods.

John Wiley & Sons, Inc., New York, 1995.

[23] S. Abarbanel and D. Gottlieb. A mathematical analysis of the PML method. Journal of

Computational Physics, 134(2):357 – 363, 1997.

[24] S. Abarbanel and D. Gottlieb. On the construction and analysis of absorbing layers in

CEM. Applied Numerical Mathematics, 27(4):331 – 340, 1998. Special Issue on Absorbing

Boundary Conditions.

214

[25] S. Abarbanel, D. Gottlieb, and J.S. Hesthaven. Long time behavior of the perfectly matched

layer equations in computational electromagnetics. Journal of Scientific Computing, 17(1-

4):405–422, 2002.

[26] K. Duru and G. Kreiss. A well-posed and discretely stable perfectly matched layer for elastic

wave equations in second order formulation. Communications in Computational Physics,

11(5):1643 – 1672, 2012.

[27] E. Becache, S. Fauqueux, and P. Joly. Stability of perfectly matched layers, group velocities

and anisotropic waves. Journal of Computational Physics, 188:399–433, 2003.

[28] F. H. Drossaert and A. Giannopoulos. A nonsplit complex frequency-shifted PML based

on recursive integration for FDTD modeling of elastic waves. Geophysics, 72(2):T9–T17,

March-April 2007.

[29] D. Komatitsch and R. Martin. An unsplit convolutional perfectly matched layer improved at

grazing incidence for the seismic wave equation. Geophysics, 72(5):SM155–SM167, September-

October 2007.

[30] T. Wang and X. Tang. Finite-difference modeling of elastic wave propagation: A nonsplitting

perfectly matched layer approach. Geophysics, 68(5):1749–1755, 2003.

[31] U. Basu. Explicit finite element perfectly matched layer for transient three-dimensional

elastic waves. Int. J. Numer. Meth. Engng., 77:151–176, 2009.

[32] R. Martin, D. Komatitsch, and S. D. Gedney. A variational formulation of a stabilized unsplit

convolutional perfectly matched layer for the isotropic or anisotropic seismic wave equation.

CMES, 37(3):274–304, 2008.

215

[33] S. Kucukcoban and L. F. Kallivokas. A symmetric hybrid formulation for transient wave

simulations in PML-truncated heterogeneous media. Wave Motion, 50:57–79, 2013.

[34] F. D. Hastings, J. B. Schneider, and S. L. Broschat. Application of the perfectly matched

layer (PML) absorbing boundary condition to elastic wave propagation. J. Acoust. Soc. Am.,

100(5):3061–3069, 1996.

[35] Q. H. Liu. Perfectly matched layers for elastic waves in cylindrical and spherical coordinates.

J. Acoust. Soc. Am., 105(4):2075–2084, April 1999.

[36] E. Becache, P. Joly, and C. Tsogka. Fictitious domains, mixed finite elements and per-

fectly matched layers for 2D elastic wave propagation. J. Comput. Acoust., 9(3):1175–1202,

September 2001.

[37] G. Festa and S. Nielsen. PML absorbing boundaries. Bulletin of the Seismological Society

of America, 93(2):891–903, April 2003.

[38] D. Komatitsch and J. Tromp. A perfectly matched layer absorbing boundary condition for

the second-order seismic wave equation. Geophysical Journal International, 154:146–153,

2003.

[39] G. Cohen and S. Fauqueux. Mixed spectral finite elements for the linear elasticity system in

unbounded domains. SIAM J. Sci. Comput., 26(3):864–884, 2005.

[40] G. Festa and J.-P. Vilotte. The Newmark scheme as velocity-stress time-staggering: an

efficient PML implementation for spectral element simulations of elastodynamics. Geophys.

J. Int., 161:789–812, 2005.

216

[41] K. C. Meza-Fajardo and A. S. Papageorgiou. A nonconvolutional, split-field, perfectly

matched layer for wave propagation in isotropic and anisotropic elastic media: Stability

analysis. Bulletin of the Seismological Society of America, 98(4):1811–1836, August 2008.

[42] D. Correia and J.-M. Jin. Performance of regular PML, CFS-PML, and second-order PML for

waveguide problems. Microwave and Optical Technology Letters, 48(10):2121–2126, October

2006.

[43] M.N. Dmitriev and V.V. Lisitsa. Application of M-PML reflectionless boundary conditions

to the numerical simulation of wave propagation in anisotropic media. part I: Reflectivity.

Numerical Analysis and Applications, 4(4):271–280, 2011.

[44] M.N. Dmitriev and V.V. Lisitsa. Application of M-PML absorbing boundary conditions

to the numerical simulation of wave propagation in anisotropic media. part II: Stability.

Numerical Analysis and Applications, 5(1):36–44, 2012.

[45] K. C. Meza-Fajardo and A. S. Papageorgiou. Study of the accuracy of the multiaxial perfectly

matched layer for the elastic-wave equation. Bulletin of the Seismological Society of America,

102(6):2458–2467, December 2012.

[46] P. Ping, Y. Zhang, and Y. Xu. A multiaxial perfectly matched layer (M-PML) for the

long-time simulation of elastic wave propagation in the second-order equations. Journal of

Applied Geophysics, 101:124 – 135, 2014.

[47] C. Bunks, F. Saleck, S. Zaleski, and G. Chavent. Multiscale seismic waveform inversion.

Geophysics, 60(5):1457–1473, 1995.

[48] R. E. Plessix. Introduction: Towards a full waveform inversion. Geophysical Prospecting,

56(6):761–763, 2008.

217

[49] A. Tarantola. Inversion of seismic reflection data in the acoustic approximation. Geophysics,

49(8):1259–1266, 1984.

[50] A. Fichtner, J. Trampert, P. Cupillard, E. Saygin, T. Taymaz, Y. Capdeville, and A. Villaseor.

Multiscale full waveform inversion. Geophysical Journal International, 2013.

[51] Q. Liu and Y. J. Gu. Seismic imaging: From classical to adjoint tomography. Tectonophysics,

566–567(0):31 – 66, 2012.

[52] P. Mora. Nonlinear two-dimensional elastic inversion of multioffset seismic data. Geophysics,

52(9):1211–1228, 1987.

[53] N. Rawlinson, S. Pozgay, and S. Fishwick. Seismic tomography: A window into deep earth.

Physics of the Earth and Planetary Interiors, 178(34):101 – 135, 2010.

[54] B. Banerjee, T. F. Walsh, W. Aquino, and M. Bonnet. Large scale parameter estimation

problems in frequency-domain elastodynamics using an error in constitutive equation func-

tional. Comp. Meth. Appl. Mech. Eng., 253:60–72, 2013.

[55] A. Gholami, A. Mang, and G Biros. Image-driven parameter estimation for low grade

gliomas. arXiv, 2014.

[56] J. W. Kang and L. F. Kallivokas. The inverse medium problem in heterogeneous PML-

truncated domains using scalar probing waves. Comput. Methods Appl. Mech. Engrg.,

200:265–283, 2011.

[57] S. Kucukcoban. The inverse medium problem in PML-truncated elastic media. Ph.D. thesis,

The University of Texas at Austin, 2010.

218

[58] S.-W. Na and L. F. Kallivokas. Direct time-domain soil profile reconstruction for one-

dimensional semi-infinite domains. Soil Dynamics and Earthquake Engineering, 29:1016–

1026, 2009.

[59] K. H. Stokoe II, S. G. Wright, J. A. Bay, and J. M. Roesset. Characterization of geotechnical

sites by SASW method. In R. D. Woods, editor, Geophysical Characterization of Sites, pages

15–25. Oxford & IBH Pub. Co., New Delhi, India, 1994.

[60] U. Albocher, P. E. Barbone, M. S. Richards, A. A. Oberai, and I. Harari. Approaches to

accommodate noisy data in the direct solution of inverse problems in incompressible plane

strain elasticity. Inverse Problems in Science and Engineering, 22(8):1307–1328, 2014.

[61] I. Epanomeritakis, V. Akcelik, O. Ghattas, and J. Bielak. A Newton-CG method for large-

scale three-dimensional elastic full-waveform seismic inversion. Inverse Problems, 24(3):034015,

2008.

[62] A. A. Oberai, N. H. Gokhale, M. M. Doyley, and M. Bamber. Evaluation of the adjoint

equation based algorithm for elasticity imaging. Physics in Medicine and Biology, 49:2955–

2974, 2004.

[63] A. A. Oberai, N. H. Gokhale, and G. R. Feijo. Solution of inverse problems in elasticity

imaging using the adjoint method. Inverse Problems, 19(2):297–313, 2003.

[64] K. T. Tran and M. McVay. Site characterization using Gauss-Newton inversion of 2-D full

seismic waveform in the time domain. Soil Dynamics and Earthquake Engineering, 43(0):16

– 24, 2012.

[65] R. G. Pratt, C. Shin, and G. J. Hick. GaussNewton and full Newton methods in frequencys-

pace seismic waveform inversion. Geophysical Journal International, 133(2):341–362, 1998.

219

[66] J. Bramwell. A Discontinuous Petrov-Galerkin Method for Seismic Tomography Problems.

Ph.D. thesis, The University of Texas at Austin, 2013.

[67] A. Fathi, B. Poursartip, and L. F. Kallivokas. Time-domain hybrid formulations for wave

simulations in three-dimensional PML-truncated heterogeneous media. International Journal

for Numerical Methods in Engineering, 101(3):165–198, 2015.

[68] N. Petra and G. Stadler. Model variational inverse problems governed by partial differential

equations. ICES Report 11-05, The Institute for Computational Engineering and Sciences,

The University of Texas at Austin, 2011.

[69] A. Fathi, L. F. Kallivokas, and B. Poursartip. Full-waveform inversion in three-dimensional

PML-truncated elastic media. Computer Methods in Applied Mechanics and Engineering,

(under review), 2015.

[70] A. Fathi, L. F. Kallivokas, K. H. Stokoe II, and B. Poursartip. Three-dimensional site

characterization using full-waveform inversion: theory, computations, and field experiments.

Soil Dynamics and Earthquake Engineering, (to be submitted), 2015.

[71] S. Kucukcoban and L.F. Kallivokas. Mixed perfectly-matched-layers for direct transient

analysis in 2D elastic heterogeneous media. Computer Methods in Applied Mechanics and

Engineering, 200(1-4):57–76, 2011.

[72] L. F. Kallivokas, J. Bielak, and R. C. MacCamy. A simple impedance-infinite element for

the finite element solution of the three-dimensional wave equation in unbounded domains.

Computer Methods in Applied Mechanics and Engineering, 147:235–262, 1997.

[73] F. Brezzi. A survey of mixed finite element methods. In D.L. Dwoyer, M.Y. Hussaini, and

R.G. Voigt, editors, Finite Elements Theory and Application, pages 34–49. Springer-Verlag,

220

New York, 1988.

[74] A. Quarteroni, R. Sacco, and F. Saleri. Numerical Mathematics. Texts in Applied Mathe-

matics. Springer, Berlin Heidelberg, second edition, 2007.

[75] A. Quarteroni and A. Valli. Numerical approximation of partial differential equations.

Springer Series in Computational Mathematics. Springer, Berlin Heidelberg, 2008.

[76] H. Bao, J. Bielak, O. Ghattas, L. F. Kallivokas, D. R. O’Hallaron, J. R. Shewchuk, and

J. Xu. Large-scale simulation of elastic wave propagation in heterogeneous media on parallel

computers. Comput. Methods Appl. Mech. Engrg., 152:85–102, 1998.

[77] D. A. Kopriva. Implementing spectral methods for partial differential equations. Springer,

2009.

[78] L.C. Wilcox, G. Stadler, C. Burstedde, and O. Ghattas. A high-order discontinuous Galerkin

method for wave propagation through coupled elastic-acoustic media. Journal of Computational

Physics, 229:9373–9396, 2010.

[79] R. Matzen. An efficient finite element time-domain formulation for the elastic second-order

wave equation: A non-split complex frequency shifted convolutional PML. International

Journal for Numerical Methods in Engineering, 88(10):951–973, 2011.

[80] M. Benzi, G. H. Golub, and J. Liesen. Numerical solution of saddle point problems. Acta

Numerica, pages 1–137, 2005.

[81] S. Balay, J. Brown, K. Buschelman, V. Eijkhout, W. D. Gropp, D. Kaushik, M. G. Knepley,

L. C. McInnes, B. F. Smith, and H. Zhang. PETSc users manual. Technical Report ANL-

95/11 - Revision 3.4, Argonne National Laboratory, 2013.

221

[82] E. Bozda, J. Trampert, and J. Tromp. Misfit functions for full waveform inversion based

on instantaneous phase and envelope measurements. Geophysical Journal International,

185(2):845–870, 2011.

[83] B. Engquist and B. Froese. Application of the Wasserstein metric to seismic signals.

Communications in Mathematical Science, 12(5):979–988, 2014.

[84] A. Tikhonov. Solution of incorrectly formulated problems and the regularization method.

Soviet Math. Dokl., 5:1035/1038, 1963.

[85] L. I. Rudin, S. Osher, and E. Fatemi. Nonlinear total variation based noise removal algo-

rithms. Phys. D, 60(1-4):259–268, November 1992.

[86] F. Troltzsch. Optimal Control of Partial Differential Equations: Theory, Methods, and

Applications, volume 112 of Graduate Studies in Mathematics. American Mathematical

Society, 2010.

[87] Z. Xie, D. Komatitsch, R. Martin, and R. Matzen. Improved forward wave propagation and

adjoint-based sensitivity kernel calculations using a numerically stable finite-element PML.

Geophysical Journal International, 198(3):1714–1747, 2014.

[88] V. Monteiller, S. Chevrot, D. Komatitsch, and N. Fuji. A hybrid method to compute short-

period synthetic seismograms of teleseismic body waves in a 3-D regional model. Geophysical

Journal International, 192(1):230–247, 2013.

[89] P. Tong, C. W. Chen, D. Komatitsch, P. Basini, and Q. Liu. High-resolution seismic

array imaging based on an SEM-FK hybrid method. Geophysical Journal International,

197(1):369–395, 2014.

222

[90] P. Tong, D. Komatitsch, T. L. Tseng, S. H. Hung, C. W. Chen, P. Basini, and Q. Liu. A

3-D spectral-element and frequency-wave number hybrid method for high-resolution seismic

array imaging. Geophysical Research Letters, 2014.

[91] Y. Choi, C. Farhat, W. Murray, and M. Saunders. A practical factorization of a Schur

complement for PDE-constrained distributed optimal control. arXiv, 2013.

[92] J. Nocedal and S. Wright. Numerical Optimization. Springer Series in Operations Research.

Springer, New York, NY, second edition, 2006.

[93] C. R. Vogel. Computational Methods for Inverse Problems. Frontiers in Applied Mathemat-

ics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2002.

[94] M. D. Gunzburger. Perspectives in Flow Control and Optimization. Advances in Design and

Control. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2003.

[95] L. C. Wilcox, G. Stadler, T. Bui-Thanh, and O. Ghattas. Discretely exact derivatives for

hyperbolic PDE-constrained optimization problems discretized by the discontinuous Galerkin

method. arXiv, 2013.

[96] S. Kucukcoban and L. F. Kallivokas. Mixed perfectly-matched-layers for direct transient

analysis in 2D elastic heterogeneous media. Comput. Methods Appl. Mech. Engrg., 200:57–

76, 2011.

[97] G. Stadler. Personal communication, 2014.

[98] E. Kausel. Fundamental Solutions in Elastodynamics: a Compendium. Cambridge Univer-

sity Press, New York, NY, 2006.

[99] D. M. Boore and J. J. Bommer. Processing of strong-motion accelerograms: needs, options

and consequences. Soil Dynamics and Earthquake Engineering, 25:93–115, 2005.

223

[100] J. G. Proakis and D. K. Manolakis. Digital Signal Processing. Prentice Hall, Englewood

Cliffs, NJ, fourth edition, 2006.

[101] J. R. Krebs, J. E. Anderson, D. Hinkley, R. Neelamani, S. Lee, A. Baumstein, and M.-D. La-

casse. Fast full-wavefield seismic inversion using encoded sources. Geophysics, 74:WCC177–

WCC188, 2009.

[102] L. A. Romero, D. C. Ghiglia, C. C. Ober, and S. A. Morton. Phase encoding of shot records

in prestack migration. Geophysics, 65:426–436, 2000.

[103] G. Meles, S. Greenhalgh, J. van der Kruk, A. Green, and H. Mauerer. Taming the non-

linearity problem in GPR full-waveform inversion for high contrast media. Journal of Applied

Geophysics, 78:31–43, 2012.

[104] S. H. Joh. Advances in the data interpretation technique for spectral-analysis-of surface-waves

(SASW) measurements. Ph.D. thesis, The University of Texas at Austin, 1996.

[105] T. Lunne, P. K. Robertson, and J.J.M. Powell. Cone Penetration Testing in Geotechnical

Practice. Spon Press, London, 2002.

[106] T. Bui-Thanh, O. Ghattas, J. Martin, and G. Stadler. A computational framework for

infinite-dimensional Bayesian inverse problems part I: The linearized case, with application

to global seismic inversion. SIAM Journal on Scientific Computing, 35(6):A2494–A2523,

2013.

[107] D. Assimaki, L.F. Kallivokas, J.W. Kang, and S. Kucukcoban. Time-domain forward and

inverse modeling of lossy soils with frequency-independent Q for near-surface applications.

Soil Dynamics and Earthquake Engineering, 43:139–159, 2012.

224

[108] A. Borzi and V. Schulz. Computational Optimization of Systems Governed by Partial

Differential Equations. Computational Science and Engineering. Society for Industrial and

Applied Mathematics (SIAM), Philadelphia, PA, 2012.

225

Vita

Arash Fathi received his B.Sc. and M.Sc. degrees in Civil Engineering from Iran. He served

in Iran Army Corps of Engineers as a structural engineer for a year and a half, and has three years

of industrial experience with the dam and hydropower industry. In January 2009, he joined the

Department of Civil, Architectural and Environmental Engineering of the University of Texas at

Austin, to pursue a Ph.D. degree in Civil Engineering / Computational Mechanics, where he has

been working as a research and teaching assistant. He is a finalist of the 26th annual Robert J.

Melosh Medal competition for the best student paper on finite element analysis.

Permanent address: [email protected].

This dissertation was typeset with LATEX† by the author.

†LATEX is a document preparation system developed by Leslie Lamport as a special version ofDonald Knuth’s TEX Program.

226